Skip to main content

Full text of "Vector Calculus"

See other formats


Vector  Calculus 


Fourth  Edition 


if*  Jr  -p 


Susan  Jane  Colley 


This  page  intentionally  left  blank 


Vector  Calculus 


This  page  intentionally  left  blank 


Vector  Calculus!  4™ 


Susan  Jane  Colley 

Oberlin  College 


PEARSON 


Boston    Columbus    Indianapolis    New  York    San  Francisco    Upper  Saddle  River 
Amsterdam    Cape  Town    Dubai    London    Madrid    Milan    Munich    Paris    Montreal  Toronto 
Delhi    Mexico  City   Sao  Paulo   Sydney    Hong  Kong   Seoul    Singapore   Taipei  Tokyo 


Editor  in  Chief:  Deirdre  Lynch 

Senior  Acquisitions  Editor:  William  Hoffman 

Sponsoring  Editor:  Caroline  Celano 

Editorial  Assistant:  Brandon  Rawnsley 

Senior  Managing  Editor:  Karen  Wernholm 

Production  Project  Manager:  Beth  Houston 

Executive  Marketing  Manager:  Jeff  Weidenaar 

Marketing  Assistant:  Caitlin  Crain 

Senior  Author  Support/Technology  Specialist:  Joe  Vetere 

Rights  and  Permissions  Advisor:  Michael  Joyce 

Manufacturing  Buyer:  Debbie  Rossi 

Design  Manager:  Andrea  Nix 

Senior  Designer:  Beth  Paquin 

Production  Coordination  and  Composition:  Aptara,  Inc. 

Cover  Designer:  Suzanne  Duda 

Cover  Image:  Alessandro  Delia  Bella/AP  Images 

Many  of  the  designations  used  by  manufacturers  and  sellers  to  distinguish  their  products  are  claimed  as 
trademarks.  Where  those  designations  appear  in  this  book,  and  Pearson  Education  was  aware  of  a  trademark 
claim,  the  designations  have  been  printed  in  initial  caps  or  all  caps. 

Library  of  Congress  Cataloging-in-Publication  Data 

Colley,  Susan  Jane. 

Vector  calculus  /  Susan  Jane  Colley.  -  4th  ed. 
p.  cm. 
Includes  index. 

ISBN-13:  978-0-321-78065-2 

ISBN-10:  0-321-78065-5 
1.  Vector  analysis.     I.  Title. 
QA433.C635  2012 
515'.63-dc23 

2011022433 

Copyright  ©  2012,  2006,  2002  Pearson  Education,  Inc. 

All  rights  reserved.  No  part  of  this  publication  may  be  reproduced,  stored  in  a  retrieval  system,  or  transmitted,  in  any  form  or  by  any 
means,  electronic,  mechanical,  photocopying,  recording,  or  otherwise,  without  the  prior  written  permission  of  the  publisher.  Printed  in 
the  United  States  of  America.  For  information  on  obtaining  permission  for  use  of  material  in  this  work,  please  submit  a  written  request 
to  Pearson  Education,  Inc.,  Rights  and  Contracts  Department,  501  Boylston  Street,  Suite  900,  Boston,  MA  021 16,  fax  your  request  to 
617-671-3447,  or  e-mail  at  http://www.pearsoned.com/legal/permissions.htm. 

123456789  10— EB— 15  14  13  12  11 


PEARSON 


www.pearsonhighered.com 


ISBN  13:  978-0-321-78065-2 
ISBN  10:  0-321-78065-5 


To  Will  and  Diane, 

with  love 


About  the  Author 
Susan  Jane  Colley 


Susan  Colley  is  the  Andrew  and  Pauline  Delaney  Professor  of  Mathematics  at 
Oberlin  College  and  currently  Chair  of  the  Department,  having  also  previously 
served  as  Chair. 

She  received  S.B.  and  Ph.D.  degrees  in  mathematics  from  the  Massachusetts 
Institute  of  Technology  prior  to  joining  the  faculty  at  Oberlin  in  1983. 

Her  research  focuses  on  enumerative  problems  in  algebraic  geometry,  partic- 
ularly concerning  multiple-point  singularities  and  higher-order  contact  of  plane 
curves. 

Professor  Colley  has  published  papers  on  algebraic  geometry  and  commuta- 
tive algebra,  as  well  as  articles  on  other  mathematical  subjects.  She  has  lectured 
internationally  on  her  research  and  has  taught  a  wide  range  of  subjects  in  under- 
graduate mathematics. 

Professor  Colley  is  a  member  of  several  professional  and  honorary  societies, 
including  the  American  Mathematical  Society,  the  Mathematical  Association  of 
America,  Phi  Beta  Kappa,  and  Sigma  Xi. 


Contents 


Preface  ix 
To  the  Student:  Some  Preliminary  Notation  xv 


1        Vectors  1 

1.1  Vectors  in  Two  and  Three  Dimensions  1 

1.2  More  About  Vectors  8 

1.3  The  Dot  Product  18 

1.4  The  Cross  Product  27 

1.5  Equations  for  Planes;  Distance  Problems  40 

1.6  Some  n-dimensional  Geometry  48 

1.7  New  Coordinate  Systems  62 
True/False  Exercises  for  Chapter  1  75 
Miscellaneous  Exercises  for  Chapter  1  75 


Differentiation  in  Several  Variables  82 

2.1  Functions  of  Several  Variables;  Graphing  Surfaces  82 

2.2  Limits  97 

2.3  The  Derivative  116 

2.4  Properties;  Higher-order  Partial  Derivatives  134 

2.5  The  Chain  Rule  142 

2.6  Directional  Derivatives  and  the  Gradient  158 

2.7  Newton's  Method  (optional)  176 
True/False  Exercises  for  Chapter  2  182 
Miscellaneous  Exercises  for  Chapter  2  183 


3       Vector-Valued  Functions  189 

3.1  Parametrized  Curves  and  Kepler's  Laws  189 

3.2  Arclength  and  Differential  Geometry  202 

3.3  Vector  Fields:  An  Introduction  221 

3.4  Gradient,  Divergence,  Curl,  and  the  Del  Operator  227 
True/False  Exercises  for  Chapter  3  237 
Miscellaneous  Exercises  for  Chapter  3  237 


4        Maxima  and  Minima  in  Several  Variables  244 

4.1  Differentials  and  Taylor's  Theorem  244 

4.2  Extrema  of  Functions  263 

4.3  Lagrange  Multipliers  278 

4.4  Some  Applications  of  Extrema  293 
True/False  Exercises  for  Chapter  4  305 
Miscellaneous  Exercises  for  Chapter  4  306 


Multiple  Integration  310 

5.1  Introduction:  Areas  and  Volumes  310 

5.2  Double  Integrals  314 

5.3  Changing  the  Order  of  Integration  334 

5.4  Triple  Integrals  337 

5.5  Change  of  Variables  349 

5.6  Applications  of  Integration  373 

5.7  Numerical  Approximations  of  Multiple  Integrals  (optional)  388 
True/False  Exercises  for  Chapter  5  401 
Miscellaneous  Exercises  for  Chapter  5  403 


6        Line  Integrals  408 

6.1  Scalar  and  Vector  Line  Integrals  408 

6.2  Green's  Theorem  429 

6.3  Conservative  Vector  Fields  439 
True/False  Exercises  for  Chapter  6  450 
Miscellaneous  Exercises  for  Chapter  6  451 


7        Surface  Integrals  and  Vector  Analysis  455 

7.1  Parametrized  Surfaces  455 

7.2  Surface  Integrals  469 

7.3  Stokes's  and  Gauss's  Theorems  490 

7.4  Further  Vector  Analysis;  Maxwell's  Equations  510 
True/False  Exercises  for  Chapter  7  522 
Miscellaneous  Exercises  for  Chapter  7  523 


8       Vector  Analysis  in  Higher  Dimensions  530 

8.1  An  Introduction  to  Differential  Forms  530 

8.2  Manifolds  and  Integrals  of  k-forms  536 

8.3  The  Generalized  Stokes's  Theorem  553 
True/False  Exercises  for  Chapter  8  561 
Miscellaneous  Exercises  for  Chapter  8  561 

Suggestions  for  Further  Reading  563 

Answers  to  Selected  Exercises  565 

Index  599 


Preface 


Physical  and  natural  phenomena  depend  on  a  complex  array  of  factors.  The  sociol- 
ogist or  psychologist  who  studies  group  behavior,  the  economist  who  endeavors 
to  understand  the  vagaries  of  a  nation's  employment  cycles,  the  physicist  who 
observes  the  trajectory  of  a  particle  or  planet,  or  indeed  anyone  who  seeks  to 
understand  geometry  in  two,  three,  or  more  dimensions  recognizes  the  need  to 
analyze  changing  quantities  that  depend  on  more  than  a  single  variable.  Vec- 
tor calculus  is  the  essential  mathematical  tool  for  such  analysis.  Moreover,  it 
is  an  exciting  and  beautiful  subject  in  its  own  right,  a  true  adventure  in  many 
dimensions. 

The  only  technical  prerequisite  for  this  text,  which  is  intended  for  a 
sophomore-level  course  in  multivariable  calculus,  is  a  standard  course  in  the  cal- 
culus of  functions  of  one  variable.  In  particular,  the  necessary  matrix  arithmetic 
and  algebra  (not  linear  algebra)  are  developed  as  needed.  Although  the  mathe- 
matical background  assumed  is  not  exceptional,  the  reader  will  still  be  challenged 
in  places. 

My  own  objectives  in  writing  the  book  are  simple  ones:  to  develop  in  students 
a  sound  conceptual  grasp  of  vector  calculus  and  to  help  them  begin  the  transition 
from  first-year  calculus  to  more  advanced  technical  mathematics.  I  maintain  that 
the  first  goal  can  be  met,  at  least  in  part,  through  the  use  of  vector  and  matrix 
notation,  so  that  many  results,  especially  those  of  differential  calculus,  can  be 
stated  with  reasonable  levels  of  clarity  and  generality.  Properly  described,  results 
in  the  calculus  of  several  variables  can  look  quite  similar  to  those  of  the  calculus 
of  one  variable.  Reasoning  by  analogy  will  thus  be  an  important  pedagogical  tool. 
I  also  believe  that  a  conceptual  understanding  of  mathematics  can  be  obtained 
through  the  development  of  a  good  geometric  intuition.  Although  I  state  many 
results  in  the  case  of  n  variables  (where  n  is  arbitrary),  I  recognize  that  the  most 
important  and  motivational  examples  usually  arise  for  functions  of  two  and  three 
variables,  so  these  concrete  and  visual  situations  are  emphasized  to  explicate  the 
general  theory.  Vector  calculus  is  in  many  ways  an  ideal  subject  for  students 
to  begin  exploration  of  the  interrelations  among  analysis,  geometry,  and  matrix 
algebra. 

Multivariable  calculus,  for  many  students,  represents  the  beginning  of  signif- 
icant mathematical  maturation.  Consequently,  I  have  written  a  rather  expansive 
text  so  that  they  can  see  that  there  is  a  story  behind  the  results,  techniques,  and 
examples — that  the  subject  coheres  and  that  this  coherence  is  important  for  prob- 
lem solving.  To  indicate  some  of  the  power  of  the  methods  introduced,  a  number 
of  topics,  not  always  discussed  very  fully  in  a  first  multivariable  calculus  course, 
are  treated  here  in  some  detail: 

•  an  early  introduction  of  cylindrical  and  spherical  coordinates  (§1.7); 

•  the  use  of  vector  techniques  to  derive  Kepler's  laws  of  planetary  motion 
(§3.1); 

•  the  elementary  differential  geometry  of  curves  in  R3,  including  discussion 
of  curvature,  torsion,  and  the  Frenet-Serret  formulas  for  the  moving  frame 
(§3.2); 

•  Taylor's  formula  for  functions  of  several  variables  (§4.1); 


X  Preface 

•  the  use  of  the  Hessian  matrix  to  determine  the  nature  (as  local  extrema)  of 
critical  points  of  functions  of  n  variables  (§4.2  and  §4.3); 

•  an  extended  discussion  of  the  change  of  variables  formula  in  double  and  triple 
integrals  (§5.5); 

•  applications  of  vector  analysis  to  physics  (§7.4); 

•  an  introduction  to  differential  forms  and  the  generalized  Stokes's  theorem 
(Chapter  8). 

Included  are  a  number  of  proofs  of  important  results.  The  more  techni- 
cal proofs  are  collected  as  addenda  at  the  ends  of  the  appropriate  sections  so 
as  not  to  disrupt  the  main  conceptual  flow  and  to  allow  for  greater  flexibility 
of  use  by  the  instructor  and  student.  Nonetheless,  some  proofs  (or  sketches  of 
proofs)  embody  such  central  ideas  that  they  are  included  in  the  main  body  of  the 
text. 

New  in  the  Fourth  Edition   

I  have  retained  the  overall  structure  and  tone  of  prior  editions.  New  features  in 
this  edition  include  the  following: 

•  210  additional  exercises,  at  all  levels; 

•  a  new,  optional  section  (§5.7)  on  numerical  methods  for  approximating 
multiple  integrals; 

•  reorganization  of  the  material  on  Newton's  method  for  approximating 
solutions  to  systems  of  n  equations  in  n  unknowns  to  its  own  (optional) 
section  (§2.7); 

•  new  proofs  in  Chapter  2  of  limit  properties  (in  §2.2)  and  of  the  general 
multivariable  chain  rule  (Theorem  5.3  in  §2.5); 

•  proofs  of  both  single-variable  and  multivariable  versions  of  Taylor's  theorem 
in  §4.1; 

•  various  additional  refinements  and  clarifications  throughout  the  text, 
including  many  new  and  revised  examples  and  explanations; 

•  new  Microsoft®  PowerPoint®  files  and  Wolfram  Mathematica®  notebooks 
that  coordinate  with  the  text  and  that  instructors  may  use  in  their  teaching 
(see  "Ancillary  Materials"  below). 

How  to  Use  This  Book  

There  is  more  material  in  this  book  than  can  be  covered  comfortably  during  a  single 
semester.  Hence,  the  instructor  will  wish  to  eliminate  some  topics  or  subtopics — or 
to  abbreviate  the  rather  leisurely  presentations  of  limits  and  differentiability.  Since 
I  frequently  find  myself  without  the  time  to  treat  surface  integrals  in  detail,  I  have 
separated  all  material  concerning  parametrized  surfaces,  surface  integrals,  and 
Stokes's  and  Gauss's  theorems  (Chapter  7),  from  that  concerning  line  integrals 
and  Green's  theorem  (Chapter  6).  In  particular,  in  a  one-semester  course  for 
students  having  little  or  no  experience  with  vectors  or  matrices,  instructors  can 
probably  expect  to  cover  most  of  the  material  in  Chapters  1-6,  although  no  doubt 
it  will  be  necessary  to  omit  some  of  the  optional  subsections  and  to  downplay 


Preface  xi 


many  of  the  proofs  of  results.  A  rough  outline  for  such  a  course,  allowing  for 
some  instructor  discretion,  could  be  the  following: 

Chapter  1       8-9  lectures 


If  students  have  a  richer  background  (so  that  much  of  the  material  in  Chapter  1 
can  be  left  largely  to  them  to  read  on  their  own),  then  it  should  be  possible  to  treat 
a  good  portion  of  Chapter  7  as  well.  For  a  two-quarter  or  two-semester  course, 
it  should  be  possible  to  work  through  the  entire  book  with  reasonable  care  and 
rigor,  although  coverage  of  Chapter  8  should  depend  on  students'  exposure  to 
introductory  linear  algebra,  as  somewhat  more  sophistication  is  assumed  there. 

The  exercises  vary  from  relatively  routine  computations  to  more  challenging 
and  provocative  problems,  generally  (but  not  invariably)  increasing  in  difficulty 
within  each  section.  In  a  number  of  instances,  groups  of  problems  serve  to  intro- 
duce supplementary  topics  or  new  applications.  Each  chapter  concludes  with  a 
set  of  miscellaneous  exercises  that  both  review  and  extend  the  ideas  introduced 
in  the  chapter. 

A  word  about  the  use  of  technology.  The  text  was  written  without  reference 
to  any  particular  computer  software  or  graphing  calculator.  Most  of  the  exercises 
can  be  solved  by  hand,  although  there  is  no  reason  not  to  turn  over  some  of  the 
more  tedious  calculations  to  a  computer.  Those  exercises  that  require  a  computer 
for  computational  or  graphical  purposes  are  marked  with  the  symbol  ^  and 
should  be  amenable  to  software  such  as  Mathematical ',  Maple®,  or  MATLAB. 

Ancillary  Materials   

In  addition  to  this  text  a  Student  Solutions  Manual  is  available.  An  Instructor's 
Solutions  Manual,  containing  complete  solutions  to  all  of  the  exercises,  is 
available  to  course  instructors  from  the  Pearson  Instructor  Resource  Center 
(www.pearsonhighered.com/irc),  as  are  many  Microsoft®  PowerPoint®  files  and 
Wolfram  Mathematica®  notebooks  that  can  be  adapted  for  classroom  use.  The 
reader  can  find  errata  for  the  text  and  accompanying  solutions  manuals  at  the 
following  address: 

www.oberlin.edu/math/faculty/colley/VCErrata.html 
Acknowledgments 

I  am  very  grateful  to  many  individuals  for  sharing  with  me  their  thoughts  and  ideas 
about  multivariable  calculus.  I  would  like  to  express  particular  appreciation  to  my 
Oberlin  colleagues  (past  and  present)  Bob  Geitz,  Kevin  Hartshorn,  Michael  Henle 
(who,  among  other  things,  carefully  read  the  draft  of  Chapter  8),  Gary  Kennedy, 
Dan  King,  Greg  Quenell,  Michael  Raney,  Daniel  Steinberg,  Daniel  Styer,  Richard 
Vale,  Jim  Walsh,  and  Elizabeth  Wilmer  for  their  conversations  with  me.  I  am  also 
grateful  to  John  Alongi,  Northwestern  University;  Matthew  Conner,  University 
of  California,  Davis;  Henry  C.  King,  University  of  Maryland;  Stephen  B.  Maurer, 


Chapter  2 
Chapter  3 
Chapter  4 
Chapter  5 
Chapter  6 


9  lectures 

4-  5  lectures 

5-  6  lectures 


8  lectures 
4  lectures 


38^11  lectures 


Swarthmore  College;  Karen  Saxe,  Macalester  College;  David  Singer,  Case  West- 
ern Reserve  University;  and  Mark  R.  Treuden,  University  of  Wisconsin  at  Stevens 
Point,  for  their  helpful  comments.  Several  colleagues  reviewed  various  versions 
of  the  manuscript,  and  I  am  happy  to  acknowledge  their  efforts  and  many  fine 
suggestions.  In  particular,  for  the  first  three  editions,  I  thank  the  following  re- 
viewers: 

Raymond  J.  Cannon,  Baylor  University; 
Richard  D.  Carmichael,  Wake  Forest  University; 
Stanley  Chang,  Wellesley  College; 

Marcel  A.  F.  Deruaz,  University  of  Ottawa  (now  emeritus); 

Krzysztof  Galicki,  University  of  New  Mexico  (deceased); 

Dmitry  Gokhman,  University  of  Texas  at  San  Antonio; 

Isom  H.  Herron,  Rensselaer  Polytechnic  Institute; 

Ashwani  K.  Kapila,  Rensselaer  Polytechnic  Institute; 

Christopher  C.  Leary,  State  University  of  New  York,  College  at  Geneseo; 

David  C.  Minda,  University  of  Cincinnati; 

Jeffrey  Morgan,  University  of  Houston; 

Monika  Nitsche,  University  of  New  Mexico; 

Jeffrey  L.  Nunemacher,  Ohio  Wesleyan  University; 

Gabriel  Prajitura,  State  University  of  New  York,  College  at  Brockport; 

Florin  Pop,  Wagner  College; 

John  T.  Scheick,  The  Ohio  State  University  (now  emeritus); 

Mark  Schwartz,  Ohio  Wesleyan  University; 

Leonard  M.  Smiley,  University  of  Alaska,  Anchorage; 

Theodore  B.  Stanford,  New  Mexico  State  University; 

James  Stasheff,  University  of  North  Carolina  at  Chapel  Hill  (now  emeritus); 

Saleem  Watson,California  State  University,  Long  Beach; 

Floyd  L.  Williams,  University  of  Massachusetts,  Amherst  (now  emeritus). 

For  the  fourth  edition,  I  thank: 

Justin  Corvino,  Lafayette  College; 

Carrie  Finch,  Washington  and  Lee  University; 

Soomin  Kim,  Johns  Hopkins  University; 

Tanya  Leise,  Amherst  College; 

Bryan  Mosher,  University  of  Minnesota. 

Many  people  at  Oberlin  College  have  been  of  invaluable  assistance  through- 
out the  production  of  all  the  editions  of  Vector  Calculus.  I  would  especially  like 
to  thank  Ben  Miller  for  his  hard  work  establishing  the  format  for  the  initial  drafts 
and  Stephen  Kasperick-Postellon  for  his  manifold  contributions  to  the  typeset- 
ting, indexing,  proofreading,  and  friendly  critiquing  of  the  original  manuscript.  I 
am  very  grateful  to  Linda  Miller  and  Michael  Bastedo  for  their  numerous  typo- 
graphical contributions  and  to  Catherine  Murillo  for  her  help  with  any  number  of 
tasks.  Thanks  also  go  to  Joshua  Davis  and  Joaquin  Espinoza  Goodman  for  their 
assistance  with  proofreading.  Without  the  efforts  of  these  individuals,  this  project 
might  never  have  come  to  fruition. 

The  various  editorial  and  production  staff  members  have  been  most  kind  and 
helpful  to  me.  For  the  first  three  editions,  I  would  like  to  express  my  appreciation 
to  my  editor,  George  Lobell,  and  his  editorial  assistants  Gale  Epps,  Melanie 
Van  Benthuysen,  and  Jennifer  Urban;  to  production  editors  Nicholas  Romanelli, 
Barbara  Mack,  and  Debbie  Ryan  at  Prentice  Hall,  and  Lori  Hazzard  at  Interactive 
Composition  Corporation;  to  Ron  Weickart  and  the  staff  at  Network  Graphics 


Preface  xiii 


for  their  fine  rendering  of  the  figures,  and  to  Tom  Benfatti  of  Prentice  Hall  for 
additional  efforts  with  the  figures;  and  to  Dennis  Kletzing  for  his  careful  and 
enthusiastic  composition  work.  For  this  edition,  it  is  a  pleasure  to  acknowledge 
my  upbeat  editor,  Caroline  Celano,  and  her  assistant,  Brandon  Rawnsley;  they 
have  made  this  new  edition  fun  to  do.  In  addition,  I  am  most  grateful  to  Beth 
Houston,  my  production  manager  at  Pearson,  Jogender  Taneja  and  the  staff  at 
Aptara,  Inc.,  Donna  Mulder,  Roger  Lipsett,  and  Thomas  Wegleitner. 

Finally,  I  thank  the  many  Oberlin  students  who  had  the  patience  to  listen  to 
me  lecture  and  who  inspired  me  to  write  and  improve  this  volume. 

SJC 

sjcolley@math.oberlin.edu 


This  page  intentionally  left  blank 


To  the  Student: 
Some  Preliminary 

Notation 

Here  are  the  ideas  that  you  need  to  keep  in  mind  as  you  read  this  book  and  learn 
vector  calculus. 

Given  two  sets  A  and  B,  I  assume  that  you  are  familiar  with  the  notation 
A  U  B  for  the  union  of  A  and  B — those  elements  that  are  in  either  A  or  B  (or 
both): 

AU  B  =  {x\x  €  Aorx  €  B}. 

Similarly,  A  n  B  is  used  to  denote  the  intersection  of  A  and  B — those  elements 
that  are  in  both  A  and  B  : 

A(~)B  =  {x\x<=A  andx  e  B}. 

The  notation  A  c  B,  or  A  c  B,  indicates  that  A  is  a  subset  of  B  (possibly  empty 
or  equal  to  B). 

One-dimensional  space  (also  called  the  real  line  or  R)  is  just  a  straight  line. 
We  put  real  number  coordinates  on  this  line  by  placing  negative  numbers  on  the 
left  and  positive  numbers  on  the  right.  (See  Figure  1.) 

Two-dimensional  space,  denoted  R2,  is  the  familiar  Cartesian  plane.  If  we 
construct  two  perpendicular  lines  (the  x-  and  y-coordinate  axes),  set  the  origin 
as  the  point  of  intersection  of  the  axes,  and  establish  numerical  scales  on  these 
lines,  then  we  may  locate  a  point  in  R2  by  giving  an  ordered  pair  of  numbers  (x ,  y), 
the  coordinates  of  the  point.  Note  that  the  coordinate  axes  divide  the  plane  into 
four  quadrants.  (See  Figure  2.) 

Three-dimensional  space,  denoted  R3,  requires  three  mutually  perpendicular 
coordinate  axes  (called  the  x-,  y-  and  z-axes)  that  meet  in  a  single  point  (called 
the  origin)  in  order  to  locate  an  arbitrary  point.  Analogous  to  the  case  of  R2,  if  we 
establish  scales  on  the  axes,  then  we  can  locate  a  point  in  R3  by  giving  an  ordered 
triple  of  numbers  (x,  y,  z).  The  coordinate  axes  divide  three-dimensional  space 
into  eight  octants.  It  takes  some  practice  to  get  your  sense  of  perspective  correct 
when  sketching  points  in  R3.  (See  Figure  3.)  Sometimes  we  draw  the  coordinate 
axes  in  R3  in  different  orientations  in  order  to  get  a  better  view  of  things.  However, 
we  always  maintain  the  axes  in  a  right-handed  configuration.  This  means  that 
if  you  curl  the  fingers  of  your  right  hand  from  the  positive  x-axis  to  the  positive 
y-axis,  then  your  thumb  will  point  along  the  positive  z-axis.  (See  Figure  4.) 

Although  you  need  to  recall  particular  techniques  and  methods  from  the 
calculus  you  have  already  learned  here  are  some  of  the  more  important  concepts 
to  keep  in  mind:  Given  a  function  f(x),  the  derivative  f'(x)  is  the  limit  (if  it  exists) 
of  the  difference  quotient  of  the  function: 

f  (x  )  =  hm  . 

J  v       ft^o  h 


XVI    To  the  Student:  Some  Preliminary  Notation 


(2,4,5) 


Figure  3  Three-dimensional 
space  R3 .  Selected  points  are 
graphed. 


Figure  4  The  x-,  y-,  and  z-axes  in  R3  are  always 
drawn  in  a  right-handed  configuration. 


The  significance  of  the  derivative  /'(xo)  is  that  it  measures  the  slope  of  the  line 
tangent  to  the  graph  of  /  at  the  point  (xo,  /(xo)).  (See  Figure  5.)  The  derivative 
may  also  be  considered  to  give  the  instantaneous  rate  of  change  of  /  at  x  =  Xq. 
We  also  denote  the  derivative  f'(x)  by  df/dx. 

The  definite  integral  fb  f(x)  dx  of  /  on  the  closed  interval  [a ,  b]  is  the  limit 
(provided  it  exists)  of  the  so-called  Riemann  sums  of  /: 


Figure  5  The  derivative  f'(xo)  is 
the  slope  of  the  tangent  line  to 
y  =  f  (x)  at  (x0,  /(x0)). 


f 


f(x)dx  =  lim 

all  Ax,-*0 


£/(x*)Ax,. 


(=1 


Here  a  =  xo  <  x\  <  xi  <  •  •  •  <  x„  =  b  denotes  a  partition  of  [a,  b]  into  subin- 
tervals  [x(_i,  x,  ],  the  symbol  A  (the  length  of  the  subinterval),  and 

x*  denotes  any  point  in  [x/_i ,  x,-].  If  f(x)  >  0  on  [a,  b],  then  each  term  f(x*)Axj 
in  the  Riemann  sum  is  the  area  of  a  rectangle  related  to  the  graph  of  /.  The 
Riemann  sum  YH=i  f(x*)^-x<  mus  approximates  the  total  area  under  the  graph 
of  /  between  x  =  a  and  x  =  b.  (See  Figure  6.) 


Figure  6  If  f(x)  >  0  on  [a,  b],  then  the  Riemann  sum 
approximates  the  area  under  y  =  f(x)  by  giving  the  sum 
of  areas  of  rectangles. 


To  the  Student:  Some  Preliminary  Notation  XVII 


The  definite  integral  fb  f(x)dx,  if  it  exists,  is  taken  to  represent  the  area 
under  y  =  f(x)  between  x  =  a  and  x  =  b.  (See  Figure  7.) 

The  derivative  and  the  definite  integral  are  connected  by  an  elegant  result 
known  as  the  fundamental  theorem  of  calculus.  Let  f(x)  be  a  continuous  func- 
tion of  one  variable,  and  let  F(x)  be  such  that  F'(x)  =  f(x).  (The  function  F  is 
called  an  antiderivative  of  /.)  Then 


I 


b 

f(x)dx  =  F(b)-F(a); 


d  fx 

2.   —  /    f(t)dt  =  f{x). 

dx  J  a 


Finally,  the  end  of  an  example  is  denoted  by  the  symbol  ♦  and  the  end  of  a 
proof  by  the  symbol  ■. 


This  page  intentionally  left  blank 


Vector  Calculus 


This  page  intentionally  left  blank 


Vectors 


1.1  Vectors  in  Two  and  Three 
Dimensions 

1 .2  More  About  Vectors 

1 .3  The  Dot  Product 

1 .4  The  Cross  Product 

1.5  Equations  for  Planes; 
Distance  Problems 

1 .6  Some  n-dimensional 
Geometry 

1.7  New  Coordinate  Systems 

True/False  Exercises  for 
Chapter  1 

Miscellaneous  Exercises  for 
Chapter  1 


1 .1   Vectors  in  Two  and  Three  Dimensions 

For  your  study  of  the  calculus  of  several  variables,  the  notion  of  a  vector  is 
fundamental.  As  is  the  case  for  many  of  the  concepts  we  shall  explore,  there  are 
both  algebraic  and  geometric  points  of  view.  You  should  become  comfortable 
with  both  perspectives  in  order  to  solve  problems  effectively  and  to  build  on  your 
basic  understanding  of  the  subject. 

Vectors  in  R2  and  R3:  The  Algebraic  Notion  


DEFINITION  1.1  A  vector  in  R2  is  simply  an  ordered  pair  of  real  numbers. 
That  is,  a  vector  in  R2  may  be  written  as 

(aua2)    (e.g.,  (1,2)  or  (n,  17)). 

Similarly,  a  vector  in  R3  is  simply  an  ordered  triple  of  real  numbers.  That  is, 
a  vector  in  R3  may  be  written  as 

(a\,a2,  a3)    (e.g.,  (jt,  e,  <s/2)). 


To  emphasize  that  we  want  to  consider  the  pair  or  triple  of  numbers  as  a 
single  unit,  we  will  use  boldface  letters;  hence  a  =  (a\,  a2)  or  a  =  (ai,  a%,  a{) 
will  be  our  standard  notation  for  vectors  in  R2  or  R3 .  Whether  we  mean  that  a  is  a 
vector  in  R2  or  in  R3  will  be  clear  from  context  (or  else  won't  be  important  to  the 
discussion).  When  doing  handwritten  work,  it  is  difficult  to  "boldface"  anything, 
so  you'll  want  to  put  an  arrow  over  the  letter.  Thus,  a  will  mean  the  same  thing 
as  a.  Whatever  notation  you  decide  to  use,  it's  important  that  you  distinguish  the 
vector  a  (or  a)  from  the  single  real  number  a.  To  contrast  them  with  vectors,  we 
will  also  refer  to  single  real  numbers  as  scalars. 

In  order  to  do  anything  interesting  with  vectors,  it's  necessary  to  develop 
some  arithmetic  operations  for  working  with  them.  Before  doing  this,  however, 
we  need  to  know  when  two  vectors  are  equal. 


DEFINITION  1.2  Two  vectors  a  =  (ai,a2)  and  b  =  (bi,b2)  in  R2  are 
equal  if  their  corresponding  components  are  equal,  that  is,  if  a\  =  b\  and 
a2  =  b2.  The  same  definition  holds  for  vectors  in  R3:  a  =  (a\,  a2,  a^)  and 
b  =  (b\,  b2,b-i)  are  equal  if  their  corresponding  components  are  equal,  that 
is,  if  a\  =b\,a2  =  b2,  and  a^  =  b^. 


2       Chapter  1  ]  Vectors 

EXAMPLE  1  The  vectors  a  =  (1, 2)  and  b  =  (§,  f )  are  equal  in  R2,  but  c  = 
(1,  2,  3)  and  d  =  (2,  3,  1)  are  not  equal  in  R3.  ♦ 

Next,  we  discuss  the  operations  of  vector  addition  and  scalar  multiplication. 
We'll  do  this  by  considering  vectors  in  R3  only;  exactly  the  same  remarks  will 
hold  for  vectors  in  R2  if  we  simply  ignore  the  last  component. 


DEFINITION  1.3  (Vector  Addition)  Let  a  =  (a\,  a2,  03)  and  b  =  (b\, 
b2,b^)  be  two  vectors  in  R3.  Then  the  vector  sum  a  +  b  is  the  vector  in  R3 
obtained  via  componentwise  addition:  a  4-  b  =  (a\  +  b\,  a2  +  b2,  03  +  ^3). 


EXAMPLE  2   We  have  (0,  1,3)  +  (7,  -2,  10)  =  (7,  - 1,  13)  and  (in  R2): 

(1,  1)  +  (7T,  V2)  =  (1  +  7T,  1  +  V2).  ♦ 

Properties  of  vector  addition.  We  have 

1.  a  +  b  =  b  +  a  for  all  a,  b  in  R3  (commutativity); 

2.  a  +  (b  +  c)  =  (a  +  b)  +  c  for  all  a,  b,  c  in  R3  (associativity); 

3.  a  special  vector,  denoted  0  (and  called  the  zero  vector),  with  the  property 
that  a  +  0  =  a  for  all  a  in  R3 . 


These  three  properties  require  proofs,  which,  like  most  facts  involving  the  al- 
gebra of  vectors,  can  be  obtained  by  explicitly  writing  out  the  vector  components. 
For  example,  for  property  1,  we  have  that  if 

a  =  {a\,a2,a{)    and    b  =  {b\,  b2,  ^3), 

then 

a  +  b  =  (czi  +  b\ ,  a2  +  b2,  «3  +  h) 
=  (b\  +  c/i ,  b2  +  a2,  b-i  +  a3) 
=  b  +  a, 

since  real  number  addition  is  commutative.  For  property  3,  the  "special  vector" 
is  just  the  vector  whose  components  are  all  zero:  0  =  (0,  0,  0).  It's  then  easy  to 
check  that  property  3  holds  by  writing  out  components.  Similarly  for  property  2, 
so  we  leave  the  details  as  exercises. 


DEFINITION  1 .4  (Scalar  Multiplication)  Let  a  =  (a\ ,  a2,  03)  be  a  vec- 
tor in  R3  and  let  k  e  R  be  a  scalar  (real  number).  Then  the  scalar  prod- 
uct ka  is  the  vector  in  R3  given  by  multiplying  each  component  of  a  by 
k:  ka  =  (ka\ ,  ka2,  ka{). 


EXAMPLE  3   Ifa  =  (2,0,  V2)and^  =  7,thenyta  =  (14,  0,  7V2).  ♦ 

The  results  that  follow  are  not  difficult  to  check — just  write  out  the  vector 
components. 


1.1  |  Vectors  in  Two  and  Three  Dimensions 


Properties  of  scalar  multiplication.  For  all  vectors  a  and  b  in  R3  (or  R2) 

and  scalars  k  and  /  in  R,  we  have 

1.  (k  +  /)a  =  ka  +  la  (distributivity); 

2.  k(a  +  b)  =  ka  +  kb  (distributivity); 

3.  k(la)  =  (kl)a  =  l(ka). 


It  is  worth  remarking  that  none  of  these  definitions  or  properties  really  de- 
pends on  dimension,  that  is,  on  the  number  of  components.  Therefore  we  could 
have  introduced  the  algebraic  concept  of  a  vector  in  R"  as  an  ordered  n-tuple 
(fli,  d2, . . . ,  an)  of  real  numbers  and  defined  addition  and  scalar  multiplication 
in  a  way  analogous  to  what  we  did  for  R2  and  R3.  Think  about  what  such  a 
generalization  means.  We  will  discuss  some  of  the  technicalities  involved  in  §  1 .6. 

Vectors  in  R2  and  R3:  The  Geometric  Notion  

Although  the  algebra  of  vectors  is  certainly  important  and  you  should  become 
adept  at  working  algebraically,  the  formal  definitions  and  properties  tend  to  present 
a  rather  sterile  picture  of  vectors.  A  better  motivation  for  the  definitions  just  given 
comes  from  geometry.  We  explore  this  geometry  now.  First  of  all,  the  fact  that 
a  vector  a  in  R2  is  a  pair  of  real  numbers  (a\ ,  ai)  should  make  you  think  of  the 
coordinates  of  a  point  in  R2.  (See  Figure  1.1.)  Similarly,  if  a  e  R3,  then  a  may 
be  written  as  (a\,  02,  as),  and  this  triple  of  numbers  may  be  thought  of  as  the 
coordinates  of  a  point  in  R3.  (See  Figure  1.2.) 

All  of  this  is  fine,  but  the  results  of  performing  vector  addition  or  scalar  mul- 
tiplication don't  have  very  interesting  or  meaningful  geometric  interpretations  in 
terms  of  points.  As  we  shall  see,  it  is  better  to  visualize  a  vector  in  R2  or  R3  as  an 
arrow  that  begins  at  the  origin  and  ends  at  the  point.  (See  Figure  1.3.)  Such  a  depic- 
tion is  often  referred  to  as  the  position  vector  of  the  point  (a\ ,  ai)  or  {a\ ,  #2,  ai). 

If  you've  studied  vectors  in  physics,  you  have  heard  them  described  as  objects 
having  "magnitude  and  direction."  Figure  1 .3  demonstrates  this  concept,  provided 
that  we  take  "magnitude"  to  mean  "length  of  the  arrow"  and  "direction"  to  be  the 
orientation  or  sense  of  the  arrow.  (Note:  There  is  an  exception  to  this  approach, 
namely,  the  zero  vector.  The  zero  vector  just  sits  at  the  origin,  like  a  point,  and  has 
no  magnitude  and  therefore,  an  indeterminate  direction.  This  exception  will  not 
pose  much  difficulty.)  However,  in  physics,  one  doesn't  demand  that  all  vectors 


4       Chapter  1  |  Vectors 


be  represented  by  arrows  having  their  tails  bound  to  the  origin.  One  is  free  to 
"parallel  translate"  vectors  throughout  R2  and  R3.  That  is,  one  may  represent 
the  vector  a  =  (a\ ,  a2,  a^)  by  an  arrow  with  its  tail  at  the  origin  (and  its  head  at 
(fli,  a2,  03))  or  with  its  tail  at  any  other  point,  so  long  as  the  length  and  sense  of 
the  arrow  are  not  disturbed.  (See  Figure  1 .4.)  For  example,  if  we  wish  to  represent 
a  by  an  arrow  with  its  tail  at  the  point  (jci,  x2,  ^3),  then  the  head  of  the  arrow 
would  be  at  the  point  {x\  +  ci\ ,  x2  +  a2,  X3  +  03).  (See  Figure  1.5.) 


Figure  1 .4  Each  arrow  is  a 
parallel  translate  of  the  position 
vector  of  the  point  (ai ,  02,  03)  and 
represents  the  same  vector. 


(^i  +  flj,  x2  +  a2,  x3  +  a3) 


(xhx2,x3) 


Figure  1 .5  The  vector 

a  =  (a\ ,  ci2,  03)  represented  by  an 

arrow  with  tail  at  the  point 

(xi,x2,  x3). 


Figure  1 .6  The  vector 
a  +  b  may  be  represented 
by  an  arrow  whose  tail  is  at 
the  tail  of  a  and  whose  head 
is  at  the  head  of  b. 


With  this  geometric  description  of  vectors,  vector  addition  can  be  visualized 
in  two  ways.  The  first  is  often  referred  to  as  the  "head-to-tail"  method  for  adding 
vectors.  Draw  the  two  vectors  a  and  b  to  be  added  so  that  the  tail  of  one  of  the 
vectors,  say  b,  is  at  the  head  of  the  other.  Then  the  vector  sum  a  +  b  may  be 
represented  by  an  arrow  whose  tail  is  at  the  tail  of  a  and  whose  head  is  at  the  head 
of  b.  (See  Figure  1.6.)  Note  that  it  is  not  immediately  obvious  that  a  +  b  =  b  +  a 
from  this  construction! 

The  second  way  to  visualize  vector  addition  is  according  to  the  so-called 
parallelogram  law:  If  a  and  b  are  nonparallel  vectors  drawn  with  their  tails  ema- 
nating from  the  same  point,  then  a  +  b  may  be  represented  by  the  arrow  (with  its 
tail  at  the  common  initial  point  of  a  and  b)  that  runs  along  a  diagonal  of  the  paral- 
lelogram determined  by  a  and  b  (Figure  1.7).  The  parallelogram  law  is  completely 
consistent  with  the  head-to-tail  method.  To  see  why,  just  parallel  translate  b  to  the 
opposite  side  of  the  parallelogram.  Then  the  diagonal  just  described  is  the  result  of 
adding  a  and  (the  translate  of)  b,  using  the  head-to-tail  method.  (See  Figure  1.8.) 

We  still  should  check  that  these  geometric  constructions  agree  with  our  alge- 
braic definition.  For  simplicity,  we'll  work  in  R2.  Let  a  =  {a\ ,  a2)  and  b  =  (b\ ,  b2) 
as  usual.  Then  the  arrow  obtained  from  the  parallelogram  law  addition  of  a  and 
b  is  the  one  whose  tail  is  at  the  origin  O  and  whose  head  is  at  the  point  P  in 
Figure  1.9.  If  we  parallel  translate  b  so  that  its  tail  is  at  the  head  of  a,  then  it  is 
immediate  that  the  coordinates  of  P  must  be  (a\  +  b\,a2  +  b2),  as  desired. 

Scalar  multiplication  is  easier  to  visualize:  The  vector  ka.  may  be  represented 
by  an  arrow  whose  length  is  \k\  times  the  length  of  a  and  whose  direction  is  the 
same  as  that  of  a  when  k  >  0  and  the  opposite  when  k  <  0.  (See  Figure  1.10.) 

It  is  now  a  simple  matter  to  obtain  a  geometric  depiction  of  the  difference 
between  two  vectors.  (See  Figure  1.11.)  The  difference  a  —  b  is  nothing  more 


1.1  |  Vectors  in  Two  and  Three  Dimensions 


a  +  b  / 

/ 

/ 

i 


a  +  b 


(translated) 


Figure  1 .7  The  vector 
a  +  b  may  be  represented 
by  the  arrow  that  runs  along 
the  diagonal  of  the 
parallelogram  determined 
by  a  and  b. 


Figure  1 .8  The  equivalence  of  the 
parallelogram  law  and  the 
head-to-tail  methods  of  vector 
addition. 


Figure  1.11  The 

geometry  of  vector 
subtraction.  The  vector  c 
is  such  that  b  +  c  =  a. 
Hence,  c  =  a  —  b. 


2< 


s_  — 

b  /  / b 

V  v 

Figure  1 .9  The  point  P  has  coordinates 
(«i  +  b\,  a2  +  b2). 


Figure  1.10  Visualization  of 
scalar  multiplication. 


than  a  +  (— b)  (where  — b  means  the  scalar  —1  times  the  vector  b).  The  vector 
a  —  b  may  be  represented  by  an  arrow  pointing  from  the  head  of  b  toward  the 
head  of  a;  such  an  arrow  is  also  a  diagonal  of  the  parallelogram  determined  by  a 
and  b.  (  As  we  have  seen,  the  other  diagonal  can  be  used  to  represent  a  +  b.) 
Here  is  a  construction  that  will  be  useful  to  us  from  time  to  time. 


DEFINITION  1 .5  Given  two  points  Pi(xi ,  yi ,  zi)  and  P2(x2,  y2,  z2)  in  R3, 
the  displacement  vector  from  P\  to  P2  is 

P\A  =  (xi  -  x\,  yi  -  yi,  zi  -  zi). 


This  construction  is  not  hard  to  understand  if  we  consider  Figure  1.12.  Given 
the  points  P\  and  P2,  draw  the  corresponding  position  vectors  OP\  and  OP2- 
Then  we  see  that  P\P2  is  precisely  OP2—OP\.  An  analogous  definition  may 
~y     be  made  for  R2. 


/  In  your  study  of  the  calculus  of  one  variable,  you  no  doubt  used  the  notions  of 

derivatives  and  integrals  to  look  at  such  physical  concepts  as  velocity,  acceleration, 
Figure  1 .1 2  The  displacement  force,  etc.  The  main  drawback  of  the  work  you  did  was  that  the  techniques  involved 
vector  P\P2,  represented  by  the  allowed  you  to  study  only  rectilinear,  or  straight-line,  activity.  Intuitively,  we  all 
arrow  from  Pi  to  P2,  is  the  understand  that  motion  in  the  plane  or  in  space  is  more  complicated  than  straight- 
difference  between  the  position  line  motion.  Because  vectors  possess  direction  as  well  as  magnitude,  they  are 
vectors  of  these  two  points.  ideally  suited  for  two-  and  three-dimensional  dynamical  problems. 


Chapter  1  |  Vectors 


Figure  1.13  After  t  seconds,  the 
point  starting  at  a,  with  velocity  v, 
moves  to  a  +  t\. 


Vj  ship 
(with  respect 
to  still  water) 


,  current 


Net  velocity 


Figure  1.14  The  length  of  Vi  is 
15,  and  the  length  of  V2  is  5V2. 


For  example,  suppose  a  particle  in  space  is  at  the  point  (ai,  a.2,  a{)  (with 
respect  to  some  appropriate  coordinate  system).  Then  it  has  position  vector  a  = 
(a\,  a2,  03).  If  the  particle  travels  with  constant  velocity  v  =  (uj,  i>2,  V3)  for  / 
seconds,  then  the  particle's  displacement  from  its  original  position  is  rv,  and  its 
new  coordinate  position  is  a  +  fv.  (See  Figure  1.13.) 

EXAMPLE  4  If  a  spaceship  is  at  position  (100,  3,  700)  and  is  traveling  with 
velocity  (7,  —10,25)  (meaning  that  the  ship  travels  7  mi/sec  in  the  positive 
x-direction,  10  mi/sec  in  the  negative  y-direction,  and  25  mi/sec  in  the  positive 
z-direction),  then  after  20  seconds,  the  ship  will  be  at  position 

(100,  3,  700)  +  20(7,  -10,  25)  =  (240,  -197,  1200), 

and  the  displacement  from  the  initial  position  is  (140,  —200,  500).  ♦ 

EXAMPLE  5  The  S.S.  Calculus  is  cruising  due  south  at  a  rate  of  15  knots 
(nautical  miles  per  hour)  with  respect  to  still  water.  However,  there  is  also  a 
current  of  5^2  knots  southeast.  What  is  the  total  velocity  of  the  ship?  If  the  ship 
is  initially  at  the  origin  and  a  lobster  pot  is  at  position  (20,  —79),  will  the  ship 
collide  with  the  lobster  pot? 

Since  velocities  are  vectors,  the  total  velocity  of  the  ship  is  Vi  +  V2,  where  vi  is 
the  velocity  of  the  ship  with  respect  to  still  water  and  V2  is  the  southeast-pointing 
velocity  of  the  current.  Figure  1.14  makes  it  fairly  straightforward  to  compute 
these  velocities.  We  have  that  vi  =  (0,  —15).  Since  \2  points  southeastward,  its 
direction  must  be  along  the  line  v  =  —  x.  Therefore,  V2  can  be  written  as  V2  = 
(v,  —v),  where  v  is  a  positive  real  number.  By  the  Pythagorean  theorem,  if  the 
length  of  \2  is  5\/2,  then  we  must  have  v2  +  (— v)2  =  (5>/2)2  or  2v2  =  50,  so 
that  v  =  5.  Thus,  v2  =  (5,  —5),  and  hence,  the  net  velocity  is 

(0,  -15) +  (5,  -5)  =  (5,  -20). 
After  4  hours,  therefore,  the  ship  will  be  at  position 

(0,0) +  4(5,  -20)  =  (20,  -80) 
and  thus  will  miss  the  lobster  pot.  ♦ 

EXAMPLE  6  The  theory  behind  the  venerable  martial  art  of  judo  is  an  excel- 
lent example  of  vector  addition.  If  two  people,  one  relatively  strong  and  the  other 
relatively  weak,  have  a  shoving  match,  it  is  clear  who  will  prevail.  For  example, 
someone  pushing  one  way  with  200  lb  of  force  will  certainly  succeed  in  overpow- 
ering another  pushing  the  opposite  way  with  1 00  lb  of  force.  Indeed,  as  Figure  1.15 
shows,  the  net  force  will  be  100  lb  in  the  direction  in  which  the  stronger  person 
is  pushing. 


1001b 


2001b 


1001b 


200  lb 


2001b 


Figure  1.16  Vector  addition  in 
judo. 


Figure  1 .1 5  A  relatively  strong  person  pushing  with  a 
force  of  200  lb  can  quickly  subdue  a  relatively  weak  one 
pushing  with  only  100  lb  of  force. 

Dr.  Jigoro  Kano,  the  founder  of  judo,  realized  (though  he  never  expressed 
his  idea  in  these  terms)  that  this  sort  of  vector  addition  favors  the  strong  over  the 
weak.  However,  if  the  weaker  participant  applies  his  or  her  100  lb  of  force  in  a 
direction  only  slightly  different  from  that  of  the  stronger,  he  or  she  will  effect  a 
vector  sum  of  length  large  enough  to  surprise  the  opponent.  (See  Figure  1.16.) 


1.1  |  Exercises 


This  is  the  basis  for  essentially  all  of  the  throws  of  judo  and  why  judo  is  described 
as  the  art  of  "using  a  person's  strength  against  himself  or  herself."  In  fact,  the 
word  "judo"  means  "the  giving  way."  One  "gives  in"  to  the  strength  of  another  by 
attempting  only  to  redirect  his  or  her  force  rather  than  to  oppose  it.  ♦ 


1.1  Exercises 


1 .  Sketch  the  following  vectors  in  R2 : 

(a)  (2,1)  (b)(3,3)  (c)(-l,2) 

2.  Sketch  the  following  vectors  in  R3 : 

(a)  (1,2,3)        (b)  (-2,0,2)  (c)  (2, -3,  1) 

3.  Perform  the  indicated  algebraic  operations.  Express 
your  answers  in  the  form  of  a  single  vector  a  =  (a  \ ,  02) 
inR2. 

(a)  (3,1) +  (-1,7) 

(b)  -2(8,  12) 

(c)  (8,9)  +  3(-l,2) 

(d)  (1,  1)  + 5(2,  6) -3(10,  2) 

(e)  (8,  10)  +  3  ((8,  -2)  -  2(4,  5)) 

4.  Perform  the  indicated  algebraic  operations.  Express 
your  answers  in  the  form  of  a  single  vector  a  = 
(ai ,  «2>  ai)  in  R3- 

(a)  (2,  1,2) +  (-3,  9,  7) 

(b)  1(8,4,  1)  +  2(5,-7,  ±) 

(c)  -2  ((2,0,  l)-6(i,-4,  1)) 

5.  Graph  the  vectors  a  =  (1,  2),  b  =  (-2,  5),  and  a  + 
b  =  ( 1 ,  2)  +  (—2,  5),  using  both  the  parallelogram  law 
and  the  head-to-tail  method. 

6.  Graph  the  vectors  a  =  (3,  2)  and  b  =  (—  1,  1).  Also 
calculate  and  graph  a  —  b,  ia,  and  a  +  2b. 

7.  Let  A  be  the  point  with  coordinates  (1,0,  2),  let  B  be 
the  point  with  coordinates  (—3,  3,  1),  and  let  C  be  the 
point  with  coordinates  (2,  1,5). 

(a)  Describe  the  vectors  AB  and  BA. 

(b)  Describe  the  vectors  AC,  B~C,  and  AC  +  CB. 

(c)  Explain,  with  pictures,  why  AC  +  CB  =  AB. 

8.  Graph  (1,2,  1 )  and  (0,-2,  3),  and  calculate  and  graph 
(1,  2,  1)  +  (0,  -2,  3),  -1(1,2,  1),  and  4(1,  2,  1). 

9.  If  (-12,  9,  z)  +  (x,  7,  -3)  =  (2,  y,  5),  what  are  x,  y, 
andz? 

10.  What  is  the  length  (magnitude)  of  the  vector  (3,  1)? 
(Hint:  A  diagram  will  help.) 

11.  Sketch  the  vectors  a  =  (l,2)andb  =  (5,  10). Explain 
why  a  and  b  point  in  the  same  direction. 


12.  Sketch  the  vectors  a  =  (2, -7, 8)  and  b=(-l, 
j,— 4).  Explain  why  a  and  b  point  in  opposite 
directions. 

13.  How  would  you  add  the  vectors  (1,2,3,4)  and 
(5,-1,  2,  0)  in  R4?  What  should  2(7,  6,  -3,  1)  be?  In 
general,  suppose  that 

a  =  (a\ ,  <22,  •  ■  ■ ,  a„)    and    b  =  (bi,  b%,  ■  ■  ■ ,  bn) 

are  two  vectors  in  R"  and  k  e  R  is  a  scalar.  Then  how 
would  you  define  a  +  b  and  fca? 

14.  Find  the  displacement  vectors  from  Pi  to  P2,  where  Pi 
and  P2  are  the  points  given.  Sketch  Pi,  P2,  and  P1P2. 

(a)  Pi(l,0,2),P2(2,  1,7) 

(b)  Pi(l,6,-l),P2(0,4,2) 

(c)  Pi(0,4,2),P2(l,6,-l) 

(d)  Pi(3,  1),P2(2,-1) 

15.  Let  Pi(2,  5,  -1,  6)  and  P2(3,  1,  -2,  7)  be  two  points 
in  R4.  How  would  you  define  and  calculate  the  dis- 
placement vector  from  Pi  to  P2?  (See  Exercise  13.) 

16.  If  A  is  the  point  in  R3  with  coordinates  (2,  5,  —6)  and 
the  displacement  vector  from  A  to  a  second  point  B  is 
(12,  —3,  7),  what  are  the  coordinates  of  Bl 

1 7.  Suppose  that  you  and  your  friend  are  in  New  York  talk- 
ing on  cellular  phones.  You  inform  each  other  of  your 
own  displacement  vectors  from  the  Empire  State  Build- 
ing to  your  current  position.  Explain  how  you  can  use 
this  information  to  determine  the  displacement  vector 
from  you  to  your  friend. 

1 8.  Give  the  details  of  the  proofs  of  properties  2  and  3  of 
vector  addition  given  in  this  section. 

1 9.  Prove  the  properties  of  scalar  multiplication  given  in 
this  section. 

20.  (a)  If  a  is  a  vector  in  R2  or  R3,  what  is  0a?  Prove  your 

answer. 

(b)  If  a  is  a  vector  in  R2  or  R3,  what  is  la?  Prove  your 
answer. 

21.  (a)  Let  a  =  (2,  0)  and  b  =  (1,  1).  For  0  <  s  <  1  and 

0  <  t  <  1,  consider  the  vector  x  =  sa  +  tb.  Ex- 
plain why  the  vector  x  lies  in  the  parallelogram 


8       Chapter  1  ]  Vectors 


determined  by  a  and  b.  (Hint:  It  may  help  to  draw 
a  picture.) 

(b)  Now  suppose  that  a  =  (2,  2,  1)  and  b  =  (0,  3,  2). 
Describe  the  set  of  vectors  {x  =  sa  +  ?b  |  0  <  .y  < 
1,  0  <  t  <  1}. 

22.  Let  a  =  {a\,  ai,  03)  andb  =  (b\,  bi,  bi)  be  two  nonzero 
vectors  such  that  b  /  ka.  Use  vectors  to  describe 
the  set  of  points  inside  the  parallelogram  with  vertex 
Po(xo,  yo,  zo)  and  whose  adjacent  sides  are  parallel  to 
a  and  b  and  have  the  same  lengths  as  a  and  b.  (See 
Figure  1.17.)  (Hint:  If  P(x,  y,  z)  is  a  point  in  the  par- 
allelogram, describe  OP,  the  position  vector  of  P.) 


z 


X 


Figure  1.17  Figure  for  Exercise  22. 


23.  A  flea  falls  onto  marked  graph  paper  at  the  point  (3,2). 
She  begins  moving  from  that  point  with  velocity  vector 
v  =  (—1,  —2)  (i.e.,  she  moves  1  graph  paper  unit  per 
minute  in  the  negative  x -direction  and  2  graph  paper 
units  per  minute  in  the  negative  y-direction). 

(a)  What  is  the  speed  of  the  flea? 

(b)  Where  is  the  flea  after  3  minutes? 

(c)  How  long  does  it  take  the  flea  to  get  to  the  point 
(-4,  -12)? 

(d)  Does  the  flea  reach  the  point  (—13,  —27)?  Why  or 
why  not? 


24.  A  plane  takes  off  from  an  airport  with  velocity  vector 
(50,  100,  4).  Assume  that  the  units  are  miles  per  hour, 
that  the  positive  a- -axis  points  east,  and  that  the  positive 
y-axis  points  north. 

(a)  How  fast  is  the  plane  climbing  vertically  at  take- 
off? 

(b)  Suppose  the  airport  is  located  at  the  origin  and  a 
skyscraper  is  located  5  miles  east  and  10  miles 
north  of  the  airport.  The  skyscraper  is  1 ,250  feet  tall. 
When  will  the  plane  be  directly  over  the  building? 

(c)  When  the  plane  is  over  the  building,  how  much 
vertical  clearance  is  there? 

25.  As  mentioned  in  the  text,  physical  forces  (e.g.,  gravity) 
are  quantities  possessing  both  magnitude  and  direction 
and  therefore  can  be  represented  by  vectors .  If  an  obj  ect 
has  more  than  one  force  acting  on  it,  then  the  resul- 
tant (or  net)  force  can  be  represented  by  the  sum  of 
the  individual  force  vectors.  Suppose  that  two  forces, 
F]  =  (2,  7,  -1)  and  F2  =  (3,  -2,  5),  act  on  an  object. 

(a)  What  is  the  resultant  force  of  Fj  and  F2? 

(b)  What  force  F3  is  needed  to  counteract  these  forces 
(i.e.,  so  that  no  net  force  results  and  the  object 
remains  at  rest)? 

26.  A  50  lb  sandbag  is  suspended  by  two  ropes.  Suppose 
that  a  three-dimensional  coordinate  system  is  intro- 
duced so  that  the  sandbag  is  at  the  origin  and  the  ropes 
are  anchored  at  the  points  (0,  —2,  1)  and  (0,  2,  1). 

(a)  Assuming  that  the  force  due  to  gravity  points  par- 
allel to  the  vector  (0,  0,-1),  give  a  vector  F  that 
describes  this  gravitational  force. 

(b)  Now,  use  vectors  to  describe  the  forces  along  each 
of  the  two  ropes.  Use  symmetry  considerations  and 
draw  a  figure  of  the  situation. 

27.  A  10  lb  weight  is  suspended  in  equilibrium  by  two 
ropes.  Assume  that  the  weight  is  at  the  point  (1,  2,  3) 
in  a  three-dimensional  coordinate  system,  where  the 
positive  z-axis  points  straight  up,  perpendicular  to  the 
ground,  and  that  the  ropes  are  anchored  at  the  points 
(3,0,4)  and  (0,3,5).  Give  vectors  Fj  and  F2  that 
describe  the  forces  along  the  ropes. 


1 .2  More  About  Vectors 

The  Standard  Basis  Vectors   

In  R2,  the  vectors  i  =  (1,  0)  and  j  =  (0,  1)  play  a  special  notational  role.  Any 
vector  a=  (a\,  Oq)  may  be  written  in  terms  of  i  and  j  via  vector  addition  and 
scalar  multiplication: 

a2)  —  (fli,  0)  +  (0,  a2)  =  fli(l,  0)  +  a2(0,  1)  =  ax  i  +  a2  j. 

(It  may  be  easier  to  follow  this  argument  by  reading  it  in  reverse.)  Insofar  as  nota- 
tion goes,  the  preceding  work  simply  establishes  that  one  can  write  either  (a\ ,  a2) 


1.2  |  More  About  Vectors 


j 

i 

«2j 

----^^  a  =  a{\  +  a2 

Figure  1.18  Any  vector  in  R2  can  be  written  in  terms  of  i  and  j. 


Figure  1.19  Any  vector  in  R  can  be  written  in  terms  of  i,  j,  and  k. 


v  =  3 

Figure  1 .20  In  R2,  the  equation 
y  =  3  describes  a  line. 


or  aji  +  ci2]  to  denote  the  vector  a.  It's  your  choice  which  notation  to  use  (as  long 
as  you're  consistent),  but  the  ij-notation  is  generally  useful  for  emphasizing  the 
"vector"  nature  of  a,  while  the  coordinate  notation  is  more  useful  for  emphasizing 
the  "point"  nature  of  a  (in  the  sense  of  a's  role  as  a  possible  position  vector  of 
a  point).  Geometrically,  the  significance  of  the  standard  basis  vectors  i  and  j  is 
that  an  arbitrary  vector  a  e  R2  can  be  decomposed  pictorially  into  appropriate 
vector  components  along  the  x-  and  y-axes,  as  shown  in  Figure  1.18. 

Exactly  the  same  situation  occurs  in  R3,  except  that  we  need  three  vec- 
tors, i  =  (1 ,  0,  0),  j  =  (0,  1 ,  0),  and  k  =  (0,  0,  1),  to  form  the  standard  basis.  (See 
Figure  1.19.)  The  same  argument  as  the  one  just  given  can  be  used  to  show  that 
any  vector  a  =  (a\,a2,  a^)  may  also  be  written  as  a\  i  +  aj.  j  +  c?3  k.  We  shall 
use  both  coordinate  and  standard  basis  notation  throughout  this  text. 


Figure  1 .21  In  R3,  the  equation 
y  =  3  describes  a  plane. 


EXAMPLE  1  We  may  write  the  vector  (1, 
(7,  n,  —3)  as  7i  +  7ij  —  3k. 


-2)  as  i  —  2j  and  the  vector 

♦ 


Parametric  Equations  of  Lines   

In  R2,  we  know  that  equations  of  the  form  y  =  mx  +  b  or  Ax  +  By  =  C  describe 
straight  lines.  (See  Figure  1.20.)  Consequently,  one  might  expect  the  same  sort  of 
equation  to  define  a  line  in  R3  as  well.  Consideration  of  a  simple  example  or  two 
(such  as  in  Figure  1.21)  should  convince  you  that  a  single  such  linear  equation 
describes  a  plane,  not  a  line.  A  pair  of  simultaneous  equations  in  x,  y,  and  z  is 
required  to  define  a  line. 

We  postpone  discussing  the  derivation  of  equations  for  planes  until  §1.5  and 
concentrate  here  on  using  vectors  to  give  sets  of  parametric  equations  for  lines  in 
R2  orR3  (or  even  R"). 


1  0       Chapter  1  |  Vectors 


t  =  3n:/2 


Figure  1 .22  The  graph  of  the 
parametric  equations  x  =  2  cos  t, 
y  =  2  sin  t ,  0  <  t  <  2n . 


First,  we  remark  that  a  curve  in  the  plane  may  be  described  analytically 
by  points  (x,  y),  where  x  and  y  are  given  as  functions  of  a  third  variable  (the 
parameter)  t.  These  functions  give  rise  to  parametric  equations  for  the  curve: 

x  =  fit) 

y  =  g(t)  ' 


EXAMPLE  2   The  set  of  equations 

x  =  2cosf 


0  <  t  <  In 


y  =  2  sin  t 

describes  a  circle  of  radius  2,  since  we  may  check  that 


x2  +  y2  =  (2  cos  t)2  +  (2  sin  t)2  =  4. 


(See  Figure  1.22.) 


Figure  1 .23  The  line  /  is  the 
unique  line  passing  through  Pq  and 
parallel  to  the  vector  a. 


Parametric  equations  may  be  used  as  readily  to  describe  curves  in  R3 ;  a  curve 
in  R3  is  the  set  of  points  (x,  y,  z)  whose  coordinates  x,y,  and  z  are  each  given  by 
a  function  of  t: 

x  =  f(t) 

y  =  g(t)  ■ 

z.  =  hit) 

The  advantages  of  using  parametric  equations  are  twofold.  First,  they  offer  a 
uniform  way  of  describing  curves  in  any  number  of  dimensions.  (How  would 
you  define  parametric  equations  for  a  curve  in  R4?  In  R128?)  Second,  they  allow 
you  to  get  a  dynamic  sense  of  a  curve  if  you  consider  the  parameter  variable  t  to 
represent  time  and  imagine  that  a  particle  is  traveling  along  the  curve  with  time 
according  to  the  given  parametric  equations.  You  can  represent  this  geometrically 
by  assigning  a  "direction"  to  the  curve  to  signify  increasing  t.  Notice  the  arrow 
in  Figure  1.22. 

Now,  we  see  how  to  provide  equations  for  lines.  First,  convince  yourself  that 
a  line  in  R2  or  R3  is  uniquely  determined  by  two  pieces  of  geometric  information: 
(1)  a  vector  whose  direction  is  parallel  to  that  of  the  line  and  (2)  any  particular 
point  lying  on  the  line — see  Figure  1.23.  In  Figure  1.24,  we  seek  the  vector 

r  =  ~OP 

between  the  origin  O  and  an  arbitrary  point  P  on  the  line  /  (i.e.,  the  position 
vector  of  P(x,  y,  z))-  OP  is  the  vector  sum  of  the  position  vector  b  of  the  given 
point  P0  (i.e.,  OPq)  and  a  vector  parallel  to  a.  Any  vector  parallel  to  a  must  be  a 
scalar  multiple  of  a.  Letting  this  scalar  be  the  parameter  variable  t,  we  have 

r=0~P~  =  OP^o  +  ta, 
and  we  have  established  the  following  proposition: 

PROPOSITION  2.1  The  vector  parametric  equation  for  the  line  through  the  point 
Poib\ ,  b2,  bi),  whose  position  vector  is  OPq  =  b  =  b\i  +  b2]  +  Z?3k,  and  parallel 
to  a  =  aii  +  a2]  +  aj,k  is 


Figure  1 .24  The  graph  of  a  line 
inR3. 


r(f)  =  b  +  /a. 


(1) 


1.2  |  More  About  Vectors  11 


Expanding  formula  (1), 

r(f)  =  OP  =  b\i  +  b2j  +  &3k  +  t{a\\  +  a2\  +  o^k) 

=  (ait  +  h)i  +  (a2t  +  b2)j  +  (a3t  +  63)k. 

Next,  write  op  as  xi  +  yj  +  zk  so  that  P  has  coordinates  (x,  y,  z).  Then,  ex- 
tracting components,  we  see  that  the  coordinates  of  P  are  (a\t  +  b\,  a2t  +  fe, 
+  b$)  and  our  parametric  equations  are 

x  =  a\t  +  b\ 
■  y  =  a2t  +  b2  ,  (2) 
z  =  a3t  +  Z?3 

where  f  is  any  real  number. 

These  parametric  equations  work  just  as  well  in  R2  (if  we  ignore  the  z- 
component)  or  in  R"  where  n  is  arbitrary.  In  R",  formula  (1)  remains  valid  where 
we  take  a  =  (a,\,  a2, . . . ,  an)  and  b  =  (b\,  b2, . . . ,  b„).  The  resulting  parametric 
equations  are 

x\  =  ci\t  +  b\ 
x2  =  a2t  +  b2 

xn  —  cin  t  -\-  bn 


Figure  1.25  Finding  equations 
for  a  line  through  two  points  in 
Example  4. 


EXAMPLE  3  To  find  the  parametric  equations  of  the  line  through  (1,  —2,  3)and 
parallel  to  the  vector  n\  —  3j  +  k,  we  have  a  =  ni  —  3j  +  k  and  b  =  i  —  2j  +  3k 
so  that  formula  (1)  yields 

r(/)  =  i  -  2j  +  3k  +  t(ni  -  3j  +  k) 

=  ( 1  +  jtt)\  +  (-2  -  3r)j  +  (3  +  r)k 

The  parametric  equations  may  be  read  as 

X  =  7tt  +  1 

y  =  -3t-2  . 
vz  =  t  +  3  ♦ 

EXAMPLE  4  From  Euclidean  geometry,  two  distinct  points  determine  a  unique 
line  in  R2  or  R3 .  Let's  find  the  parametric  equations  of  the  line  through  the  points 
Pq(\,  —2,  3)  and  P\(0,  5,  —1).  The  situation  is  suggested  by  Figure  1.25.  To  use 
formula  (1),  we  need  to  find  a  vector  a  parallel  to  the  desired  line.  The  vector  with 
tail  at  Pq  and  head  at  P\  is  such  a  vector.  That  is,  we  may  use  for  a  the  vector 


PoP^  =  (0-1,5-  (-2),  -1  -  3)  =  -i  +  7j  -  4k. 

For  b,  the  position  vector  of  a  particular  point  on  the  line,  we  have  the  choice 
of  taking  either  b  =  i  —  2j  +  3k  or  b  =  5j  —  k.  Hence,  the  equations  in  (2)  yield 
parametric  equations 


x  =  1  -  t 
y  =  -2  +  It 
z  =  3  -  At 


or 


x  =  —  t 
y  =  5  +  It 

z  =  -1  -At 


Chapter  1  |  Vectors 


In  general,  given  two  arbitrary  points 

Po(a\,  a2,  a3)    and    P\(b\,  b2,  fc3), 
the  line  joining  them  has  vector  parametric  equation 


r(t)  =  OP0  +  tP0Pi. 
Equation  (3)  gives  parametric  equations 

x  =  a\  +  (b\  —  a\)t 
y  =  a2  +  (b2  -  a2)t  . 
Z  =  a3  +  (b2  -  ai)t 

Alternatively,  in  place  of  equation  (3),  we  could  use  the  vector  equation 


(3) 


(4) 


r(r)  =  OP\  +  tP0P\ 


or  perhaps 


r(r)  =  OPx  +  tP\P0 


(5) 


(6) 


each  of  which  gives  rise  to  somewhat  different  sets  of  parametric  equations.  Again, 

we  refer  you  to  Figure  1 .25  for  an  understanding  of  the  vector  geometry  involved. 

Example  4  brings  up  an  important  point,  namely,  that  parametric  equations 

for  a  line  (or,  more  generally,  for  any  curve)  are  never  unique.  In  fact,  the  two 

sets  of  equations  calculated  in  Example  4  are  by  no  means  the  only  ones;  we 
 > 

could  have  taken  a  =  Pi  Pq  =  i  —  7j  +  4k  or  any  nonzero  scalar  multiple  of 
P^p\  for  a. 

If  parametric  equations  are  not  determined  uniquely,  then  how  can  you  check 
your  work?  In  general,  this  is  not  so  easy  to  do,  but  in  the  case  of  lines,  there  are 
two  approaches  to  take.  One  is  to  produce  two  points  that  lie  on  the  line  specified 
by  the  first  set  of  parametric  equations  and  see  that  these  points  lie  on  the  line 
given  by  the  second  set  of  parametric  equations.  The  other  approach  is  to  use  the 
parametric  equations  to  find  what  is  called  the  symmetric  form  of  a  line  in  R3 . 
From  the  equations  in  (2),  assuming  that  each  a,  is  nonzero,  one  can  eliminate 
the  parameter  variable  t  in  each  equation  to  obtain: 

x  —  b\ 


t  = 


t  = 


y  -  b2 
a2 

z  -  h 
a  3 


The  symmetric  form  is 


1.2  |  More  About  Vectors  13 


In  Example  4,  the  two  sets  of  parametric  equations  give  rise  to  corresponding 
symmetric  forms 

x  —  1      y  +  2      z  —  3        ,     x       y  —  5      z  + 1 


■1 


and 


■1 


It's  not  difficult  to  see  that  adding  1  to  each  "side"  of  the  second  symmetric  form 
yields  the  first  one.  In  general,  symmetric  forms  for  lines  can  differ  only  by  a 
constant  term  or  constant  scalar  multiples  (or  both). 

The  symmetric  form  is  really  a  set  of  two  simultaneous  equations  in  R3.  For 
example,  the  information  in  (7)  can  also  be  written  as 

b\      y  -  b2 


x 

fll 
x  —  b\ 


y 

fl3 


This  illustrates  that  we  require  two  "scalar"  equations  in  x,  y,  and  z  to  describe  a 
line  in  R3,  although  a  single  vector  parametric  equation,  formula  (1),  is  sufficient. 

The  next  two  examples  illustrate  how  to  use  parametric  equations  for  lines  to 
identify  the  intersection  of  a  line  and  a  plane  or  of  two  lines. 

EXAMPLE  5   We  find  where  the  line  with  parametric  equations 

x  =  t  +  5 


y 


-2t  -  4 
3t  +  7 


intersects  the  plane  3x  +  2y  —  lz  =  2. 

To  locate  the  point  of  intersection,  we  must  find  what  value  of  the  parameter  t 
gives  a  point  on  the  line  that  also  lies  in  the  plane.  This  is  readily  accomplished  by 
substituting  the  parametric  values  for  x,  y,  and  z  from  the  line  into  the  equation 
for  the  plane 

3(r  +  5)  +  2(-2t  -  4)  -  7(3f  +  7)  =  2.  (8) 

Solving  equation  (8)  for  t,  we  find  that  t  =  —2.  Setting  t  equal  to  —2  in  the 
parametric  equations  for  the  line  yields  the  point  (3,0,  1),  which,  indeed  lies  in 
the  plane  as  well.  ♦ 


EXAMPLE  6   We  determine  whether  and  where  the  two  lines 

3t  -  3 
and     \  y  =  t 


x  —  t  +  1 
y  =  5t  +  6 
z  =  -2t 


x 

y 

z  =  t+  1 


intersect. 

The  lines  intersect  provided  that  there  is  a  specific  value  t\  for  the  parameter 
of  the  first  line  and  a  value  h  for  the  parameter  of  the  second  line  that  generate  the 
same  point.  In  other  words,  we  must  be  able  to  find  t\  and  t2  so  that,  by  equating 
the  respective  parametric  expressions  for  x,  y,  and  z„  we  have 


h  +  1   =  3f2  - 
5fi  +  6  =  t2 
-2t{     =  t2  +  1 


3 


(9) 


1  4       Chapter  1  |  Vectors 


The  last  two  equations  of  (9)  yield 

t2  =  5?i  +  6  =  -2*i  -  1     =>     fi  =  —  1. 


Using  t\  =  —  1  in  the  second  equation  of  (9),  we  find  that  t2  =  1 .  Note  that  the 
values  t\  =  —  1  and  ?2  =  1  also  satisfy  the  first  equation  of  (9);  therefore,  we  have 
solved  the  system.  Setting  t  =  —  1  in  the  set  of  parametric  equations  for  the  first 
line  gives  the  desired  intersection  point,  namely,  (0,  1,2).  ♦ 


Parametric  Equations  in  General  

Vector  geometry  makes  it  relatively  easy  to  find  parametric  equations  for  a  variety 
of  curves.  We  provide  two  examples. 

EXAMPLE  7  If  a  wheel  rolls  along  a  flat  surface  without  slipping,  a  point  on 
the  rim  of  the  wheel  traces  a  curve  called  a  cycloid,  as  shown  in  Figure  1.26. 


Figure  1 .27  The  result  of  the 
wheel  in  Figure  1 .26  rolling 
through  a  central  angle  of  t . 


D   

3x12  -  A 

A 

Figure  1 .28  AP  with  its  tail  at 
the  origin. 


Figure  1 .26  The  graph  of  a  cycloid. 

Suppose  that  the  wheel  has  radius  a  and  that  coordinates  in  R2  are  chosen  so  that 
the  point  of  interest  on  the  wheel  is  initially  at  the  origin.  After  the  wheel  has 
rolled  through  a  central  angle  of  t  radians,  the  situation  is  as  shown  in  Figure  1.27. 
We  seek  the  vector  OP,  the  position  vector  of  P,  in  terms  of  the  parameter  t. 
Evidently,  OP  =  OA  +  AP,  where  the  point  A  is  the  center  of  the  wheel.  The 
vector  O  A  is  not  difficult  to  determine.  Its  j-component  must  be  a ,  since  the  center 
of  the  wheel  does  not  vary  vertically.  Its  i-component  must  equal  the  distance  the 
wheel  has  rolled;  if  t  is  measured  in  radians,  then  this  distance  is  at,  the  length 
of  the  arc  of  the  circle  having  central  angle  t.  Hence,  OA  =  ati  +  aj. 

The  value  of  vector  methods  becomes  apparent  when  we  determine  A~P. 
Parallel  translate  the  picture  so  that  A~P  has  its  tail  at  the  origin,  as  in  Figure  1 .28. 
From  the  parametric  equations  of  a  circle  of  radius  a, 


aP  = 


a  cos 


3tt 

T 


+  a  sin 


3tt 
T 


?  I  j  =  —a  sin  t\  —  a  cos  t  j, 


from  the  addition  formulas  for  sine  and  cosine.  We  conclude  that 

OP  =  0~A  +  A~P  =  (ati  +  aj)  +  (—a  sin  ti  —  a  cos  t\) 
=  a(t  —  sin  t  )i  +  a(l  —  cos  t)j, 


1.2  |  More  About  Vectors  15 


so  the  parametric  equations  are 

x  =  a(t  —  sinf) 

y  =  a(l  —  cos  t)  ♦ 

EXAMPLE  8  If  you  unwind  adhesive  tape  from  a  nonrotating  circular  tape 
dispenser  so  that  the  unwound  tape  is  held  taut  and  tangent  to  the  dispenser  roll, 
then  the  end  of  the  tape  traces  a  curve  called  the  involute  of  the  circle.  Let's 
find  the  parametric  equations  for  this  curve,  assuming  that  the  dispensing  roll 
has  constant  radius  a  and  is  centered  at  the  origin.  (As  more  and  more  tape  is 
unwound,  the  radius  of  the  roll  will,  of  course,  decrease.  We'll  assume  that  little 
enough  tape  is  unwound  so  that  the  radius  of  the  roll  remains  constant.) 

Considering  Figure  1 .29,  we  see  that  the  position  vector  O  P  of  the  desired 
point  P  is  the  vector  sum  OB  +  BP.  To  determine  O B  and  BP,  we  use  the  angle 
9  between  the  positive  x-axis  and  OB  as  our  parameter.  Since  B  is  a  point  on  the 
circle, 

OB  =  a  cos#  i  +  a  sin9  j. 


Unwound 
\  tape 

B      /  Involute 

\  0 

J  (a,  0) 

Figure  1 .29  Unwinding  tape,  as 
in  Example  8.  The  point  P 
describes  a  curve  known  as  the 
involute  of  the  circle. 


Figure  1 .30  The  vector  BP  must 
make  an  angle  of  9  —  n/2  with  the 
positive  x-axis. 


To  find  the  vector  B  P,  parallel  translate  it  so  that  its  tail  is  at  the  origin.  Figure  1 .30 
shows  that  B~P 's  length  must  be  a9,  the  amount  of  unwound  tape,  and  its  direction 
must  be  such  that  it  makes  an  angle  of  0  —  n/2  with  the  positive  x-axis.  From  our 
experience  with  circular  geometry  and,  perhaps,  polar  coordinates,  we  see  that 
BP  is  described  by 


iO  cos     —  |)  i  +  ad  sin  (0  —  — ^  j  =  a6  sin0  i  —  a9  cosO  j. 

OP  =  ~0%  +  B~^  =  a(cos6>  +6>sin0)i  +  a(sin6»  -  6>cos6>)j. 

Ix  =  a(cos9  +  0  sin#) 
y  =  a(sin  0  —  9  cos  0) 

are  the  parametric  equations  of  the  involute,  whose  graph  is  pictured  in 
Figure  1.31.  ♦ 


BP 
Hence, 

So 


Chapter  1  |  Vectors 

1.2  Exercises 


In  Exercises  1-5,  write  the  given  vector  by  using  the  standard 
basis  vectors  for  R2  and  R3. 


1.  (2,4)  2.  (9,-6) 

4.  (-1,2,5)         5.  (2,4,0) 


3.  (3,;r, -7) 


In  Exercises  6—10,  write  the  given  vector  without  using  the 
standard  basis  notation. 

6.  i  +  j  -3k 

7.  9i-2j  +  V2k 

8.  -3(2i-7k) 

9.  jri  —  j  (Consider  this  to  be  a  vector  in  R2.) 

10.  jri  —  j  (Consider  this  to  be  a  vector  in  R3.) 

11.  Letai  =(1,  1)  and  a2  =  (1,-1). 

(a)  Write  the  vector  b  =  (3, 1)  as  ciai  +  C2a2,  where 
c\  and  C2  are  appropriate  scalars. 

(b)  Repeat  part  (a)  for  the  vector  b  =  (3,  —  5). 

(c)  Show  that  any  vector  b  =  {b\,  bj)  in  R2  may  be 
written  in  the  form  c^i  +  C2&2  for  appropriate 
choices  of  the  scalars  c\,  cj,  (This  shows  that  ai 
and  a2  form  a  basis  for  R2  that  can  be  used  instead 
of  i  and  j.) 

12.  Leta!=(l,0, -l),a2  =  (0,  1,  0),  and  a3  =(1,  1,-1). 

(a)  Find  scalars  c\,  c%,  C3,  so  as  to  write  the  vector 
b  =  (5,  6,  —5)  as  Qai  +  C2a2  +  c^. 

(b)  Try  to  repeat  part  (a)  for  the  vector  b  =  (2,  3,  4). 
What  happens? 

(c)  Can  the  vectors  ai,  a2,  a3  be  used  as  a  basis  for 
R3,  instead  of  i,  j,  k?  Why  or  why  not? 

In  Exercises  13—18,  give  a  set  of  parametric  equations  for  the 
lines  so  described. 

1 3.  The  line  in  R3  through  the  point  (2,  —  1 ,  5)  that  is  par- 
allel to  the  vector  i  +  3j  —  6k. 

14.  The  line  in  R3  through  the  point  (12,  -2,  0)  that  is 
parallel  to  the  vector  5i  —  12j  +  k. 

1 5.  The  line  in  R2  through  the  point  (2,  —  1 )  that  is  parallel 
to  the  vector  i  —  7j. 

16.  The  line  in  R3  through  the  points  (2,1,2)  and 
(3,-1,5). 

17.  The  line  in  R3  through  the  points  (1,4,5)  and 
(2,4,-1). 

18.  The  line  in  R2  through  the  points  (8,  5)  and  (1,  7). 

1 9.  Write  a  set  of  parametric  equations  for  the  line  in  R4 
through  the  point  (1,  2,  0,  4)  and  parallel  to  the  vector 
(-2,5,3,7). 


20.  Write  a  set  of  parametric  equations  for  the  line  in 
R5  through  the  points  (9,^,-1,5,2)  and  (-1,1, 
V2,7,  1). 

21 .  (a)  Write  a  set  of  parametric  equations  for  the  line  in 

R3  through  the  point  (—  1 ,  7,  3)  and  parallel  to  the 
vector  2i  —  j  +  5k. 

(b)  Write  a  set  of  parametric  equations  for  the  line 
through  the  points  (5,-3,  4)  and  (0,  1,  9). 

(c)  Write  different  (but  equally  correct)  sets  of  equa- 
tions for  parts  (a)  and  (b). 

(d)  Find  the  symmetric  forms  of  your  answers  in 
(a)-(c). 

22.  Give  a  symmetric  form  for  the  line  having  parametric 
equations  x  =  5  —  2t,  y  =  3t  +  1,  z  =  6t  —  4. 

23.  Give  a  symmetric  form  for  the  line  having  parametric 
equations  x  =  t  +  1,  y  =  3t  —  9,  z  =  6  —  8f . 

24.  A  certain  line  in  R3  has  symmetric  form 

x  -  2      v-3  z+l 


5  -2  4 

Write  a  set  of  parametric  equations  for  this  line. 

25.  Give  a  set  of  parametric  equations  for  the  line  with 
symmetric  form 

x  +  5     y-1  z+10 


3  7-2 
26.  Are  the  two  lines  with  symmetric  forms 
x  -  I      y  +  2  z+l 


and 


x  -  4      y-1      z  +  5 


10        -5  : 
the  same?  Why  or  why  not? 
27.  Show  that  the  two  sets  of  equations 


z         x  +  1      y  +  6      z  +  5 

-  and   =  '-  =  

5  _6        -14  -10 


x  —  2      y  —  1 

3     =  7 
actually  represent  the  same  line  in  R3 . 

28.  Determine  whether  the  two  lines  / 1  and  I2  defined  by 
the  sets  of  parametric  equations  h:  x  =  2t  —  5,  y  = 
3t  +  2,  z  =  1  -  6t,  and  l2:  x  =  1  -  2t,  y  =  11  -  3f, 
z  =  6t  —  17  are  the  same.  (Hint:  First  find  two  points 
on  l\  and  then  see  if  those  points  lie  on  I2.) 

29.  Do  the  parametric  equations  l\\  x  =  3t  +  2,  y  = 
t  -  7,  z  =  5t+l,  and  l2:  x  =  6t  -  1,  y  =  2t  -  8, 
z  =  10?  —  3  describe  the  same  line?  Why  or  why  not? 


1.2  |  Exercises 


30.  Do  the  parametric  equations  x  =  3f3  +  7,  y  =  2  —  f3, 
Z  =  5/3  +  1  determine  a  line?  Why  or  why  not? 

31.  Do  the  parametric  equations  x  =  5t2  —  1,  y  =  2t2  + 
3,  z  =  1  —  t2  determine  a  line?  Explain. 

32.  A  bird  is  flying  along  the  straight-line  path  x  =  2t  +  7, 
y  =  t  —  2,  z=  1  —  3f,  where/ is  measured  in  minutes. 

(a)  Where  is  the  bird  initially  (at  t  =  0)?  Where  is  the 
bird  3  minutes  later? 

(b)  Give  a  vector  that  is  parallel  to  the  bird's  path. 

(c)  When  does  the  bird  reach  the  point  ( y ,  | ,  —  y  )? 

(d)  Does  the  bird  reach  (17,  4,  -14)? 

33.  Find  where  the  line  x  =  3t  —  5,  y  =  2  —  t,  z  =  6t  in- 
tersects the  plane  x  +  3y  —  z  =  19. 

34.  Where  does  the  line  x  =  1  —  At,  y  =  t  —  3/2,  z  = 
2t  +  1  intersect  the  plane  5x  —  2y  +  z  =  1? 

35.  Find  the  points  of  intersection  of  the  line  x  =  2t  —  3, 
y  =  3t  +2,  z  =  5  —  t  with  each  of  the  coordinate 
planes  x  =  0,  y  =  0,  and  z  =  0. 

36.  Show  that  the  line  x  =  5  —  t,  y  =  2t  —  7,  z  =  t  —  3  is 
contained  in  the  plane  having  equation  2x  —  y  +  4z  =  5. 

37.  Does  the  line  x  =  5  —  t,  y  =  2t  —  3,  z  =  It  +  1  inter- 
sect the  plane  x  —  3y  +  z  =  1?  Why? 

38.  Find  where  the  line  having  symmetric  form 

x  —  3_y  +  2_z 
6     ~     3     ~  5 

intersects  the  plane  with  equation  2x  —  5v  +  3z  +  8  =  0. 

39.  Show  that  the  line  with  symmetric  form 

x -3  z+2 

^r  =  y-5  =  — 

lies  entirely  in  the  plane  3x  +  3 y  +  z  =  22. 

40.  Does  the  line  with  symmetric  form 

x+4  _ y-2 _ z-\ 
3     ~    -1    ~  -9 

intersect  the  plane  2x  —  3y  +  z  =  7? 

41 .  Let  a,b,cbe  nonzero  constants.  Show  that  the  line  with 
parametric  equations  x  =  at  +  a,  y  =  b,  z  =  ct  +  c 
lies  on  the  surface  with  equation  x2/a2  +  y2/b2  — 
Z2/c2  =  1. 

42.  Find  the  point  of  intersection  of  the  two  lines  l\ :  x  = 
2t  +  3,  y  =  3t  +  3,  z  =  2t  +  1  and  l2:x  =  15  -  It, 
y  =  t-2,z  =  3t-l. 

43.  Do  the  lines  lx:x  =  2t+\,  y  =  -3t,  z  =  t-\ 
and  l2:x  =  3t  +  1,  y  =  t  +  5,  z  =  7  -t  intersect? 
Explain  your  answer. 


44.  (a)  Find  the  distance  from  the  point  (—2,  1,  5)  to  any 

point  on  the  line  x  =  3t  —  5,  y  =  1  —  t,z  =  4t  +  7. 
(Your  answer  should  be  in  terms  of  the  param- 
eter t.) 

(b)  Now  find  the  distance  between  the  point  (—2,  1 ,  5) 
and  the  line  x  =  3t  -  5,  y  =  1  -  t,  z  =  At  +  7. 
(The  distance  between  a  point  and  a  line  is  the  dis- 
tance between  the  given  point  and  the  closest  point 
on  the  line.) 

45.  (a)  Describe  the  curve  given  parametrically  by 

x  =  2cos3r     „  27r 
0  <  t  <  — . 
y  =  2  sin3r  3 

What  happens  if  we  allow  t  to  vary  between  0 
and  27r? 

(b)  Describe  the  curve  given  parametrically  by 

x  =  5cos3f     „  27r 
0  <  t  <  — . 
y  =  5  sin3r  3 

(c)  Describe  the  curve  given  parametrically  by 

{x  =  5  sin3t      „  27r 
0  <  t  <  — . 
y  =  5  cos3f  3 

(d)  Describe  the  curve  given  parametrically  by 

x  =  5  cos  3t     „  27r 
0  <  t  <  — . 
y  =  3  sin3r  3 

46.  Suppose  that  a  bicycle  wheel  of  radius  a  rolls  along  a 
fiat  surface  without  slipping.  If  a  reflector  is  attached 
to  a  spoke  of  the  wheel  at  a  distance  b  from  the  center, 
the  resulting  curve  traced  by  the  reflector  is  called  a 
curtate  cycloid.  One  such  cycloid  appears  in  Fig- 
ure 1.32,  where  a  =  3  and  b  =  2. 


y 


1  1 

1 

2ji 

i 

4?r 

Figure  1 .32  A  curtate  cycloid. 


Using  vector  methods  or  otherwise,  find  a  set  of  para- 
metric equations  for  the  curtate  cycloid.  Figure  1.33 
should  help.  (Take  a  low  point  of  the  cycloid  to  lie 


1  8       Chapter  1  |  Vectors 


Figure  1 .34  A  prolate  cycloid. 


Figure  1 .33  The  point  P  traces  a 
curtate  cycloid. 

on  the  v-axis.)  There  is  no  theoretical  reason  that  the 
cycloid  just  described  cannot  have  a  <  b,  although  in 
such  case  the  bicycle-wheel-reflector  application  is  no 
longer  relevant.  (When  a  <  b,  the  parametrized  curve 
that  results  is  called  a  prolate  cycloid.)  Your  paramet- 
ric equations  should  be  such  that  the  constants  a  and  b 
can  be  chosen  independently  of  one  another.  An  exam- 
ple of  a  prolate  cycloid,  with  a  =  2  and  b  =  4,  is  shown 
in  Figure  1.34.  Try  to  think  of  a  physical  situation  in 
which  such  a  curve  would  arise. 


47.  Egbert  is  unwinding  tape  from  a  circular  dispenser  of 
radius  a  by  holding  the  tape  taut  and  perpendicular  to 
the  dispenser.  Find  a  set  of  parametric  equations  for 
the  path  traced  by  the  end  of  the  tape  (the  point  P  in 
Figure  1 .35)  as  Egbert  unwinds  the  tape.  Use  the  angle 
0  between  OP  and  the  positive  x -axis  for  parameter. 
Assume  that  little  enough  tape  is  unwound  so  that  the 
radius  of  the  dispenser  remains  constant. 


p 

/ 

u 

Figure  1.35  Figure  for  Exercise  47. 


1.3   The  Dot  Product 

When  we  introduced  the  arithmetic  notions  of  vector  addition  and  scalar  mul- 
tiplication, you  may  well  have  wondered  why  the  product  of  two  vectors  was 
not  defined.  You  might  think  that  "vector  multiplication"  should  be  defined  in  a 
manner  analogous  to  the  way  we  defined  vector  addition  (i.e.,  by  componentwise 
multiplication).  However,  such  a  definition  is  not  very  useful.  Instead  we  shall 
define  and  use  two  different  concepts  of  a  product  of  two  vectors:  (1)  the  Eu- 
clidean inner  product,  or  "dot"  product,  which  may  be  defined  for  two  vectors  in 
R"  (where  n  is  arbitrary)  and  (2)  the  "cross"  or  vector  product,  which  is  defined 
only  for  vectors  in  R3 . 


1.3  |  The  Dot  Product  19 
The  Dot  Product  of  Two  Vectors   


DEFINITION  3.1  Let  a  =  (a\,a2,  a3)  and  b  =  (bu  b2,  b3)  be  two  vectors 
in  R3.  The  dot  (or  inner  or  scalar)  product  of  a  and  b,  denoted  a  •  b,  is 

a  •  b  =  a\b\  +  a2b2  +  03^3  ■ 

In  R2,  the  analogous  definition  is 

a  •  b  =  a\b\  +  a2b2, 

where  a  =  (a\ ,  a2)  and  b  =  (b\ ,  b2). 


EXAMPLE  1    In  R3,  we  have 

(1,  -2,  5)  •  (2,  1,  3)  =  (1)(2)  +  (-2X1)  +  (5X3)  =  15. 
(3i  +  2j  -  k)  •  (i  -  2k)  =  (3X1)  +  (2X0)  +  (-1X-2)  =  5.  # 

In  accordance  with  its  name,  the  dot — or  scalar — product  takes  two  vectors 
and  produces  a  single  real  number  (not  a  vector). 

The  following  facts  are  consequences  of  Definition  3.1: 


Properties  of  dot  products.  If  a,  b,  and  c  are  any  vectors  in  R3  (or  R2)  and 
k  e  R  is  any  scalar,  then 

1.  a  •  a  >  0,  and  a  •  a  =  0  if  and  only  if  a  =  0; 

2.  a-b  =  ba; 

3.  a  •  (b  +  c)  =  a  •  b  +  a  •  c; 

4.  (jfca)  •  b  =  jfc(a  •  b)  =  a  •  (kb). 


Proof  of  Property  1  If  a  =  (a\,  a2,  a^),  then  we  have 

a  •  a  =  a\ci\  +  a2a2  +  aj,a^  =  a\  +  a\+  a\. 

This  last  expression  is  evidently  nonnegative,  since  it  is  a  sum  of  squares  of  real 
numbers.  Moreover,  such  an  expression  is  zero  exactly  when  each  of  the  terms  is 
zero,  that  is,  if  and  only  if  a\  =  a2  =     =  0.  ■ 

We  leave  the  proofs  of  properties  2,  3,  and  4  as  exercises. 

Thus  far,  we  have  introduced  the  dot  product  of  two  vectors  as  a  purely 
algebraic  construction.  It  is  the  geometric  interpretation  of  the  definition  that  is 
really  interesting.  To  establish  this  interpretation,  we  begin  with  the  following: 


DEFINITION  3.2  If  a  =  (a\,a2,  a3),  then  the  length  of  a  (also  called  the 
norm  or  magnitude),  denoted  ||a||,  is  ^/a2x  +  a2  +  a2. 


Chapter  1  |  Vectors 


a 


Figure  1.36  The  dot 

product  of  a  and  b  is 
llall  llbll  cosd. 


a 


Figure  1 .37  The  vector 
triangle  used  in  the  proof 
of  Theorem  3.3. 


The  motivation  for  this  definition  is  evident  if  we  draw  a  as  the  position  vector  of 
the  point  (a\ ,  a%,  a$).  Then  the  length  of  the  arrow  from  the  origin  to  {a\ ,  02,  03)  is 

V(ai  -  0)2  +  (a2  "  0)2  +  (a3  -  0)2, 

as  given  by  the  distance  formula,  which  is  nothing  more  than  an  extension  of  the 
Pythagorean  theorem  in  the  plane.  As  we  just  saw,  a  ■  a  =  a\  +  a\  +  a\,  and  we 
have 

||a||  =  v/aT a 

or,  equivalently, 


a  -  a  =  ||a||2.  (1) 


Now  we're  ready  to  state  the  main  result  concerning  the  geometry  of  the  dot 
product.  If  a  and  b  are  two  nonzero  vectors  in  R3  (or  R2)  drawn  with  their  tails  at 
the  same  point,  let  9,  where  0  <  9  <  it,  be  the  angle  between  a  and  b.  If  either  a 
or  b  is  the  zero  vector,  then  9  is  indeterminate  (i.e.,  can  be  any  angle). 

THEOREM  3.3    If  a  and  b  are  any  two  vectors  in  either  R2  or  R3,  then 

a-b=  ||a||  ||b||  cos (9. 

(See  Figure  1.36.) 


PROOF  If  either  a  or  b  is  the  zero  vector,  say  a,  then  a  =  (0,  0,  0)  and  so 

a.b  =  (0)(&i)  +  (0)(&2)  +  (0)0>3)  =  0. 
Also,  ||  a||  =  0  in  this  case,  so  the  formula  in  Theorem  3.3  holds.  In  this  case,  the 
angle  6  is  indeterminate. 

Now  suppose  that  neither  a  nor  b  is  the  zero  vector.  Let  c  =  b  —  a.  Then 
we  may  apply  the  law  of  cosines  to  the  triangle  whose  sides  are  a,  b,  and  c 
(Figure  1.37)  to  obtain 

Hell2  =  ||a||2  +  ||b||2  -2||a||  ||b||  cose. 

Thus, 

2 1| a||  ||b||  cose  =  ||a||2  +  ||b||2  -  ||c||2  =  a  •  a  +  b  •  b  -  c  •  c,  (2) 
from  equation  (1).  Now,  use  the  properties  of  the  dot  product.  Since  c  =  b  —  a, 
c  •  c  =  (b  -  a)  •  (b  —  a) 

=  (b  -  a)  •  b  -  (b  -  a)  •  a 

=  b-  b-  a-  b-  b-  a  +  a-a,  (3) 

by  properties  3  and  4  of  the  dot  product.  If  we  use  equation  (3)  to  substitute  for 
c  •  c  in  equation  (2),  then 

2 1|  a  ||  ||b||  cos6»  =  a-a  +  b-b-(b-b-a-b-b-a  +  a-a) 

=  a  •  b  +  b  •  a 

=  2a-b, 

by  property  2  of  the  dot  product.  By  canceling  the  factor  of  2  on  both  sides,  the 
desired  result  is  obtained.  ■ 


1.3  |  The  Dot  Product  21 


Figure  1 .38  An  object 
sliding  down  a  ramp. 
The  force  due  to  gravity 
is  downward,  but  the 
direction  of  travel  of  the 
object  is  inclined  30°  to 
the  horizontal. 


Angles  Between  Vectors   

Theorem  3.3  may  be  used  to  find  the  angle  between  two  nonzero  vectors  a  and 
b — just  solve  for  6  in  the  formula  in  Theorem  3.3  to  obtain 


The  use  of  the  inverse  cosine  is  unambiguous,  since  we  take  0  <  6  <  it  when 
defining  angles  between  vectors. 


EXAMPLE  2   If  a  =  i  +  j  and  b  =  j  -  k,  then  formula  (4)  gives 


=  cos 


-i  (i  +  j)-(j 


=  cos 


1 


=  COS 


-1 


1 


7T 

3' 


+  jllllj-k||  (V2-V2) 

If  a  and  b  are  nonzero,  then  Theorem  3.3  implies 

cos  6  =  0    if  and  only  if   a  •  b  =  0. 

We  have  cos  6>  =  Ojust  incased  =  jt/2.  (Remember  our  restriction  on  0.)  Hence, 
it  makes  sense  for  us  to  call  a  and  b  perpendicular  (or  orthogonal)  when  a  •  b  = 
0.  If  either  a  or  b  is  the  zero  vector,  then  we  cannot  use  formula  (4),  and  the  angle 
9  is  undefined.  Nonetheless,  since  a  •  b  =  0  if  a  or  b  is  0,  we  adopt  the  standard 
convention  and  say  that  the  zero  vector  is  perpendicular  to  every  vector. 

EXAMPLE  3   The  vector  i  +  j  is  orthogonal  to  the  vector  i  —  j  +  k,  since 

(i  +  j)  •  (i  -  j  +  k)  =  (1X1)  +  (IX- 1)  +  (0X1)  =  0-  ♦ 


Vector  Projections   

Suppose  that  a  2  kg  object  is  sliding  down  a  ramp  having  a  30°  incline  with  the 
horizontal  as  in  Figure  1.38.  If  we  neglect  friction,  the  only  force  acting  on  the 
object  is  gravity.  What  is  the  component  of  the  gravitational  force  in  the  direction 
of  motion  of  the  object? 

To  answer  questions  of  this  nature,  we  need  to  find  the  projection  of  one  vector 
on  another.  The  general  idea  is  as  follows:  Given  two  nonzero  vectors  a  and  b, 
imagine  dropping  a  perpendicular  line  from  the  head  of  b  to  the  line  through  a. 
Then  the  projection  of  b  onto  a,  denoted  projab,  is  the  vector  represented  by  the 
arrow  in  Figure  1.39. 


projab  projab 
Figure  1 .39  Projection  of  the  vector  b  onto  the  vector  a. 


22       Chapter  1  |  Vectors 


Given  this  intuitive  understanding  of  the  projection,  we  find  a  precise  formula 
for  it.  Recall  that  a  vector  is  determined  by  magnitude  (length)  and  direction.  It 
follows  by  definition  that  the  direction  of  projab  is  either  the  same  as  that  of  a, 
or  opposite  to  a  if  the  angle  9  between  a  and  b  is  more  than  jt/2.  Trigonometry 
then  tells  us 

,     m      II  Projab  || 
cos#  =  -  -. 


(The  absolute  value  sign  around  cos#  is  needed  in  case  jt/2  <  0  <  jt.)  Hence, 
with  a  bit  of  algebra,  we  have 

II  -kII  mi  .I  I  m  ||a||  ||b|||cos0|  |a-b| 
||proJab||  =  ||b|||cos0|  =  =  — 

by  Theorem  3.3.  Thus,  we  know  the  magnitude  and  direction  of  projab.  To  obtain 
a  compact  formula  for  projab,  note  the  following: 

PROPOSITION  3.4  Let  k  be  any  scalar  and  a  any  vector.  Then 

1.  II ka\\  =  \k\  ||a||. 

2.  A  unit  vector  (i.e.,  a  vector  of  length  1)  in  the  direction  of  a  nonzero  vector 
a  is  given  by  a/||a||. 


PROOF  Part  1  is  left  as  an  exercise.  (Write  out  ka  and  ||&a||  in  terms  of  compo- 
nents.) For  part  2,  we  must  check  that  the  length  of  a/||a||  is  1: 


a 

1 

 a 

Ilall 

Ilall 

|a||  =  1, 


by  part  1  (since  l/||a||  is  a  positive  scalar). 

Now  projab  is  a  vector  of  length  a  •  b|, 

|a-b|\  a 
Tal 


projab  =  ± 


|  a ||  in  the  "±a-direction."  That  is, 
_±  ||a||  ||b||  |cosfj|  a 


length  of 
projab 


unit  vector 
in  direction  of  a 


Note  that  the  angle  9  keeps  track  of  the  appropriate  sign  of  projab;  that  is,  when 
0  <  6  <  jt/2,  cos  0  is  positive  and  projab  points  in  the  direction  of  a,  and  when 
7r/2  <  8  <  Tt,  cos  9  is  negative  and  projab  points  in  the  direction  opposite  to  that 
of  a.  Thus,  we  can  eliminate  both  the  ±  sign  and  the  absolute  value,  and  we  find 
that 

||a||  ||b||  cos<9  a  a-b 

proi„b  =  =  r  a 

Ilall        ||a||  ||a||2 

by  Theorem  3.3,  so  that 


by  equation  (1).  Formula  (5)  is  concise  and  not  difficult  to  remember. 


1.3  |  The  Dot  Product 


Figure  1 .40  The  2  kg 

object  sliding  down  a 
ramp  in  Example  4. 


EXAMPLE  4  To  answer  the  question  posed  at  the  beginning  of  this  subsection, 
we  need  to  calculate  projaF,  where  F  is  the  gravitational  force  vector  and  a  points 
along  the  ramp  as  shown  in  Figure  1 .40.  We  have  a  coordinate  situation  as  shown 
in  Figure  1.41.  From  trigonometric  considerations,  we  must  have  a  =  a\\  +  a^j 
such  that  flj  =  —  ||a||  cos  30°  and  a2  =  —  ||a||  sin 30°.  Since  we  are  really  only 
interested  in  the  direction  of  a,  there  is  no  loss  in  assuming  that  a  is  a  unit  vector. 
Thus, 

a  =  -  cos 30° i- sin 30°  j  =  -^i-  \\. 


al^b° 

F  =  -mgj  =  -19.6j 

Figure  1 .41  The  vectors  a  and  F  in  Example  4, 
realized  in  a  coordinate  system. 


Taking  g  =  9.8  m/sec  ,  we  have  F 
implies 


-2gj  =  — 19. 6j.  Therefore,  formula  (5) 


.  a-F 

projaF  =  |           |  a  = 


a  •  a 


(-  |j)-(-19.6j)  /   V3.  O 


27 

«  -8.49i  -  4.9 j, 
and  the  component  of  F  in  this  direction  is 

||projaF||  =  ||-8.49i-4.9j||  =9.8N.  ♦ 

Unit  vectors — that  is,  vectors  of  length  1 — are  important  in  that  they  capture 
the  idea  of  direction  (since  they  all  have  the  same  length).  Part  2  of  Proposition 
3.4  shows  that  every  nonzero  vector  a  can  have  its  length  adjusted  to  give  a  unit 
vector  u  =  a/ 1|  a||  that  points  in  the  same  direction  as  a.  This  operation  is  referred 
to  as  normalization  of  the  vector  a. 

EXAMPLE  5  A  fluid  is  flowing  across  a  plane  surface  with  uniform  velocity 
vector  v.  If  n  is  a  unit  vector  perpendicular  to  the  plane  surface,  let's  find  (in  terms 
of  v  and  n)  the  volume  of  the  fluid  that  passes  through  a  unit  area  of  the  plane  in 
unit  time.  (See  Figure  1.42.) 


24       Chapter  1  |  Vectors 


Figure  1 .42  Fluid  flowing  across 
a  plane  surface. 


Base  has  area  1 

Figure  1 .43  After  one  unit  of  time,  the  fluid  passing 
across  a  square  will  have  filled  the  box. 


First,  imagine  one  unit  of  time  has  elapsed.  Then  over  a  unit  area  of  the  plane 
(say  over  a  unit  square),  the  fluid  will  have  filled  a  "box"  as  in  Figure  1.43.  The 
box  may  be  represented  by  a  parallelepiped  (a  three-dimensional  analogue  of  a 
parallelogram).  The  volume  we  seek  is  the  volume  of  this  parallelepiped  and  is 

Volume  =  (area  of  base)(height). 

The  area  of  the  base  is  1  unit  by  construction.  The  height  is  given  by  ||projnv||. 
From  formula  (5), 

/n-v\ 

Pr°Jnv  =  (  )  n  =  (n  •  v)n, 

vn  •  n/ 

since  n  •  n  =  ||n||2  =  1.  Hence, 

||projnv||  =  ||(n-v)n||  =  |n-v|  ||n||  =  |n-v|, 
by  part  1  of  Proposition  3.4.  ♦ 


Vector  Proofs   

We  conclude  this  section  with  two  illustrations  of  how  wonderfully  well  vectors 
are  suited  to  providing  elegant  proofs  of  geometric  results. 

EXAMPLE  6  In  an  arbitrary  triangle,  show  that  the  line  segment  joining  the 
midpoints  of  two  sides  is  parallel  to  and  has  half  the  length  of  the  third  side.  (See 
Figure  1.44.)  In  other  words,  if  M\  is  the  midpoint  of  side  AB  and  M2  is  the 
midpoint  of  side  AC,  we  wish  to  show  that  M\  M2  is  parallel  to  BC  and  has  half 
its  length. 


A  A 


Figure  1 .44  In  triangle  ABC,  Figure  1 .45  The  vector  version 

Mi  M2  is  parallel  to  B  C  and  has  of  triangle  ABC  in  Example  6. 

half  its  length. 

For  a  vector  proof,  we  use  the  diagram  in  Figure  1 .45,  a  slightly  modified  ver- 
sion of  Figure  1 .44.  The  midpoint  conditions  translate  to  the  following  statements 
about  vectors: 

AA?i  =  ±A$,       A~\f2  =  \At. 


1.3  |  The  Dot  Product 


Now, 

M1M2  =  AM2  -  ~AM\  =  \AC  -  \lh  =  \{AC  -  AB)  =  \~B~£. 

But  M\ M2  =  \BC  is  precisely  what  we  wish  to  prove:  To  say  M\M2  is  a  scalar 

times  BC  means  that  the  two  vectors  are  parallel.  Moreover,  from  part  1  of 
Proposition  3.4, 

\\Mjt2\\  =  \\\BC\\  =  I||fi£||, 
so  that  the  length  condition  also  holds.  ♦ 

EXAMPLE  7  Show  that  every  angle  inscribed  in  a  semicircle  is  a  right  angle, 
as  suggested  by  Figure  1 .46. 


Figure  1.46  Every  angle  Figure  1.47  a  and  b  are  "radius 

inscribed  in  a  semicircle  is  a  vectors." 
right  angle. 


To  prove  this  remark,  we'll  make  use  of  Figure  1 .47,  where  a  and  b  are  "radius 
vectors"  with  tails  at  the  center  of  the  circle.  We  need  only  show  that  a  —  b  (a 
vector  along  one  ray  of  the  angle  in  question)  is  perpendicular  to  —  a  —  b  (a  vector 
along  the  other  ray).  In  other  words,  we  wish  to  show  that 

(a  -  b)  •  (-a  -  b)  =  0. 

We  have 

(a  -  b)  •  (-a  -  b)  =  (-l)(a  -  b)  •  (a  +  b), 
by  property  4  of  dot  products, 

=  (-l)((a-b)-a  +  (a-b)-b) 
=  (-l)(a  a-b  a  +  a  b-b  b) 
=  (-l)(||a||2-  ||b||2), 

by  properties  2  and  4, 

=  0, 

since  both  a  and  b  are  radius  vectors  (and  therefore  have  the  same  length,  namely, 
the  radius  of  the  circle).  ♦ 

Vector  proofs  as  in  Examples  6  and  7  are  elegant  and  sometimes  allow  you  to 
write  shorter  and  more  direct  proofs  than  those  from  your  high  school  geometry 
days. 


26       Chapter  1  [  Vectors 

1.3  Exercises 


Compute  a  •  b,  ||a||,  ||b||  for  the  vectors  listed  in  Exercises  1—6. 


1. 

a  = 

(l,5),b  =  (-2,3) 

2. 

a  = 

(4,-l),b  =  (±,2) 

3. 

a  = 

(-1,0,  7),  b  =  (2,  4,  -6) 

4. 

a  = 

(2,  1,  0),  b  =  (1,  -2,  3) 

5. 

a  = 

4i  -  3j  +  k,  b  =  i  +  j  +  k 

6. 

a  = 

i  +  2j-k,b  =  -3j  +  2k 

In  Exercises  7-11,  find  the  angle  between  each  of  the  pairs  of 
vectors. 


7. 

a  = 

V3i  +  j,b  =  — a/3  i  +  j 

8. 

a 

(-1,2),  b  =  (3,  1) 

9. 

a  = 

i  +  j,b  =  i  +  j  +  k 

10. 

a  = 

i  +  j  -  k,  b  =  -i  +  2j  +  2k 

11. 

a  = 

(l,-2,3),b  =  (3,-6, -5) 

In  Exercises  12—16,  calculate  projab. 

12.  a  =  i  + j,  b  =  2i  +  3j  -  k 

13.  a  =  (i  +  j)/V2,  b  =  2i  +  3j-k 

14.  a  =  5k,  b  =  i-j  +  2k 

15.  a=  -3k,b  =  i-j  +  2k 

16.  a  =  i  +  j  +  2k,  b  =  2i  -  4j  +  k 

1 7.  Give  a  unit  vector  that  points  in  the  same  direction  as 
the  vector  2i  —  j  +  k. 

1 8.  Give  a  unit  vector  that  points  in  the  direction  opposite 
to  the  vector  — i  +  2k. 

1 9.  Give  a  vector  of  length  3  that  points  in  the  same  direc- 
tion as  the  vector  i  +  j  —  k. 

20.  Find  three  nonparallel  vectors  that  are  perpendicular 
to  i  -  j  +  k. 

21.  Is  it  ever  the  case  that  projab  =  projba?  If  so,  under 
what  conditions? 

22.  Prove  properties  2,  3,  and  4  of  dot  products. 

23.  Prove  part  1  of  Proposition  3.4. 

24.  Suppose  that  a  force  F  =  i  —  2j  is  acting  on  an  object 
moving  parallel  to  the  vector  a  =  4i  +  j.  Decompose  F 
into  a  sum  of  vectors  Fi  and  F2,  where  F;  points  along 
the  direction  of  motion  and  F2  is  perpendicular  to  the 
direction  of  motion.  (Hint:  A  diagram  may  help.) 


25.  In  physics,  when  a  constant  force  acts  on  an  object 
as  the  object  is  displaced,  the  work  done  by  the  force 
is  the  product  of  the  length  of  the  displacement  and 
the  component  of  the  force  in  the  direction  of  the  dis- 
placement. Figure  1 .48  depicts  an  object  acted  upon  by 
a  constant  force  F,  which  displaces  it  from  the  point  P 
to  the  point  Q.  Let  6  denote  the  angle  between  F  and 
the  direction  of  displacement. 

(a)  Show  that  the  work  done  by  F  is  determined  by  the 
formula  F- Pg. 

(b)  Find  the  work  done  by  the  (constant)  force  F  = 
i  +  5j  +  2k  in  moving  a  particle  from  the  point 
(1,  -1,  1)  to  the  point  (2,  0,  -1). 


F 

6L 


>Q 


Component  of  F  in  direction 
of  displacement 

Figure  1.48  A  constant  force  F 
displaces  the  object  from  P  to  Q.  (See 
Exercise  25.) 

26.  A  refrigerator  is  dragged  12  ft  across  a  smooth  floor 
using  a  rope  and  60  lb  of  force  directed  along  the  rope. 
How  much  work  is  done  if  the  rope  makes  a  20°  angle 
with  the  horizontal? 

27.  How  much  work  is  done  in  pushing  a  handtruck  loaded 
with  500  lb  of  bananas  40  ft  up  a  ramp  inclined  30° 
from  horizontal? 

Let  a  be  a  nonzero  vector  in  R3.  The  direction  cosines  of  a  are 

the  three  numbers  cos  a,  cos  ft,  cos  y  determined  by  the  angles 
a,  fi,  y  between  a  and,  respectively,  the  positive  x-,  y-,  and 
Z-axes.  In  Exercises  28  and  29,  find  the  direction  cosines  of  the 
given  vectors. 


28.  a 

29.  a 


i  +  2j-k 
3i  +  4k 


30.  If  a  =  flii  +  02)  +  03k,  give  expressions  for  the  direc- 
tion cosines  of  a  in  terms  of  the  components  of  a. 

31.  Let  A,  P,  and  C  denote  the  vertices  of  a  triangle.  Let 
0  <  r  <  1 .  If  Pi  is  the  point  on  AS  located  r  times  the 
distance  from  A  to  5  and  Pi  is  the  point  on  AC  located 
r  times  the  distance  from  A  to  C,  use  vectors  to  show 
that  Pi  P2  is  parallel  to  PC  and  has  r  times  the  length 
of  PC.  (This  result  generalizes  that  of  Example  6  of 
this  section.) 


1.4  |  The  Cross  Product 


32.  Let  A,  B,  C,  and  D  be  four  points  in  R3  such  that  no 
three  of  them  lie  on  a  line.  Then  ABCD  is  a  quadri- 
lateral, though  not  necessarily  one  that  lies  in  a  plane. 
Denote  the  midpoints  of  the  four  sides  of  ABCD  by 
Mi,  M%,  M3,  and  M4.  Use  vectors  to  show  that,  amaz- 
ingly, M1M2M2M4  is  always  a  parallelogram. 

33.  Use  vectors  to  show  that  the  diagonals  of  a  parallel- 
ogram have  the  same  length  if  and  only  if  the  paral- 
lelogram is  a  rectangle.  (Hint:  Let  a  and  b  be  vectors 
along  two  sides  of  the  parallelogram.  Express  vectors 
running  along  the  diagonals  in  terms  of  a  and  b.  See 
Figure  1.49.) 


a 


Figure  1.49  Diagram  for  Exercise  33. 

34.  Using  vectors,  prove  that  the  diagonals  of  a  parallelo- 
gram are  perpendicular  if  and  only  if  the  parallelogram 
is  a  rhombus.  (Note:  A  rhombus  is  a  parallelogram 
whose  four  sides  all  have  the  same  length.) 

35.  This  problem  concerns  three  circles  of  equal  radius  r 
that  intersect  in  a  single  point  O.  (See  Figure  1.50.) 

(a)  Let  W\,  W2,  and  W3  denote  the  centers  of  the 

three  circles  and  let  w,-  =  OWi  for  i  =  1,  2,  3. 
Similarly,  let  A,  B,  and  C  denote  the  remaining 

intersection  points  of  the  circles  and  set  a  =  OA, 
b  =  OB,  and  c  =  OC.  By  numbering  the  centers 
of  the  circles  appropriately,  write  a,  b,  and  c  in 
terms  of  wi ,  W2,  and  W3 . 

(b)  Show  that  A,  B,  and  C  lie  on  a  circle  of  the 
same  radius  r  as  the  three  given  circles.  (Hint: 


The  center  of  the  circle  is  at  the  point  P,  where 

OP  =  Wl  +  W2  +  W3 .) 


Figure  1.50  Two  examples 
of  three  circles  of  equal  radius 
intersecting  in  a  single 
point  O.  (See  Exercise  35.) 


(c)  Show  that  O  is  the  orthocenter  of  triangle  ABC. 
(The  orthocenter  of  a  triangle  is  the  common  in- 
tersection point  of  the  altitudes  perpendicular  to 
the  edges.) 

36.  (a)  Show  that  the  vectors  [|b[|a+  [|a||b  and  ||b||a  — 
||  a  ||  b  are  orthogonal. 

(b)  Show  that  ||b||a  +  ||a||b  bisects  the  angle  between 
a  and  b. 


1 .4  The  Cross  Product 

The  cross  product  of  two  vectors  in  R3  is  an  "honest"  product  in  the  sense  that  it 
takes  two  vectors  and  produces  a  third  one.  However,  the  cross  product  possesses 
some  curious  properties  (not  the  least  of  which  is  that  it  cannot  be  defined  for 
vectors  in  R2  without  first  embedding  them  in  R3  in  some  way)  making  it  less 
"natural"  than  may  at  first  seem  to  be  the  case. 

When  we  defined  the  concepts  of  vector  addition,  scalar  multiplication,  and 
the  dot  product,  we  did  so  algebraically  (i.e.,  by  a  formula  in  the  vector  compo- 
nents) and  then  saw  what  these  definitions  meant  geometrically.  In  contrast,  we 
will  define  the  cross  product  first  geometrically,  and  then  deduce  an  algebraic  for- 
mula for  it.  This  technique  is  more  convenient,  since  the  coordinate  formulation 


28       Chapter  1  |  Vectors 


a 

Figure  1 .51  The  area  of  this 
parallelogram  is  ||a||  ||b||  sin#. 


Figure  1 .53  i  x  j  =  k. 


Figure  1 .54  A  mnemonic  for 
finding  the  cross  product  of  the 
unit  basis  vectors. 


is  fairly  complicated  (although  we  will  find  a  way  to  organize  it  so  as  to  make  it 
easier  to  remember). 


The  Cross  Product  of  Two  Vectors  in  R3 


DEFINITION  4.1  Let  a  and  b  be  two  vectors  in  R3  (not  R2).  The  cross 
product  (or  vector  product)  of  a  and  b,  denoted  a  x  b,  is  the  vector  whose 
length  and  direction  are  given  as  follows: 

•  The  length  of  a  x  b  is  the  area  of  the  parallelogram  spanned  by  a  and  b 
or  is  zero  if  either  a  is  parallel  to  b  or  if  a  or  b  is  0.  Alternatively,  the 
following  formula  holds: 

||axb||  =  ||a||  ||b||  sin  6, 

where  0  is  the  angle  between  a  and  b.  (See  Figure  1.51.) 

•  The  direction  of  a  x  b  is  such  that  a  x  b  is  perpendicular  to  both  a  and 
b  (when  both  a  and  b  are  nonzero)  and  is  taken  so  that  the  ordered  triple 
(a,  b,  a  x  b)  is  a  right-handed  set  of  vectors,  as  shown  in  Figure  1.52. 
(If  either  a  or  b  is  0,  or  if  a  is  parallel  to  b,  then  a  x  b  =  0  from  the 
aforementioned  length  condition.) 

By  saying  that  (a,  b,  a  x  b)  is  right-handed  we  mean  that  if  you  let  the  fingers 
of  your  right  hand  curl  from  a  toward  b,  then  your  thumb  will  point  in  the 
direction  of  a  x  b. 


EXAMPLE  1  Let's  compute  the  cross  product  of  the  standard  basis  vectors 
for  R3.  First  consider  i  x  j  as  shown  in  Figure  1.53.  The  vectors  i  and  j  deter- 
mine a  square  of  unit  area.  Thus,  ||i  x  j||  =  1.  Any  vector  perpendicular  to  both 
i  and  j  must  be  perpendicular  to  the  plane  in  which  i  and  j  lie.  Hence,  i  x  j 
must  point  in  the  direction  of  ±k.  The  "right-hand  rule"  implies  that  i  x  j  must 
point  in  the  positive  k  direction.  Since  ||k||  =  1,  we  conclude  that  i  x  j  =  k.  The 
same  argument  establishes  that  j  x  k  =  i  and  k  x  i  =  j.  To  remember  these  ba- 
sic equations,  you  can  draw  i,  j,  and  k  in  a  circle,  as  in  Figure  1.54.  Then  the 
relations 


ixj  =  k,       jxk  =  i,       kxi  =  j  (1) 


may  be  read  from  the  circle  by  beginning  at  any  vector  and  then  proceeding 
clockwise.  ♦ 


Properties  of  the  Cross  Product;  Coordinate  Formula  

Example  1  demonstrates  that  the  calculation  of  cross  products  from  the  geometric 
definition  is  not  entirely  routine.  What  we  really  need  is  a  coordinate  formula, 
analogous  to  that  for  the  dot  product  or  for  vector  projections,  which  is  not  difficult 
to  obtain. 


1.4  I  The  Cross  Product  29 
From  our  Definition  4.1,  it  is  possible  to  establish  the  following: 


Properties  of  the  Cross  Product.  Let  a,  b,  and  c  be  any  three  vectors  in  R3 
and  let  k  e  R  be  any  scalar.  Then 

1.  axb  =  —  bxa  (anticommutativity); 

2.  ax(b  +  c)  =  axb  +  axc  (distributivity); 

3.  (a  +  b)xc  =  axc  +  bxc  (distributivity); 

4.  k(a  x  b)  =  (fca)  x  b  =  a  x  (kb). 


We  provide  proofs  of  these  properties  at  the  end  of  the  section,  although  you 
might  give  some  thought  now  as  to  why  they  hold.  It's  worth  remarking  that  these 
properties  are  entirely  reasonable,  ones  that  we  would  certainly  want  a  product  to 
have.  However,  you  should  be  clear  about  the  fact  that  the  cross  product  fails  to 
satisfy  other  properties  that  you  might  also  consider  to  be  eminently  reasonable. 
In  particular,  since  property  1  holds,  we  see  that  axb/bxain  general  (i.e., 
the  cross  product  is  not  commutative).  Consequently,  be  very  careful  about  the 
order  in  which  you  write  cross  products.  Another  property  that  the  cross  product 
does  not  possess  is  associativity.  That  is, 

a  x  (b  x  c)  /  (a  x  b)  x  c, 

in  general.  For  example,  let  a  =  b  =  i  and  c  =  j.  Then 

i  x  (i  x  j)  =  i  x  k  =  -k  x  i  =  -j, 

from  properties  1  and  4,  but  (ixi)xj  =  0xj  =  0^— j.  (The  equation 
i  x  i  =  0  holds  because  i  is,  of  course,  parallel  to  i.)  Make  sure  that  you  do 
not  try  to  use  an  associative  law  when  working  problems. 

We  now  have  the  tools  for  producing  a  coordinate  formula  for  the  cross 
product.  Let  a  =  a\i  +  a2\  +  a^k  and  b  =  b\\  +  b2\  +  b^k.  Then 

a  x  b  =  (a\i  +  a2\  +  a^k)  x  (b\\  +  b2\  +  b^k) 

=  (aii  +  a2\  +  03k)  x  bii  +  (flii  +  a2\  +  a3k)  x  b2j 

+  (aii  +  a2)  +  03k)  x  Z?3k, 

by  property  2, 

=  aib\\  x  i  +  a2b\j  x  i  +  c^ik  x  i  +  a\b2\  x  j  +  a2b2\  x  j 
+  a^b2k  x  j  +  aib^i  x  k  +  a2b^i  x  k  +  a^b^k  x  k, 

by  properties  3  and  4.  These  nine  terms  may  look  rather  formidable  at  first,  but 
we  can  simplify  by  means  of  the  formulas  in  (1),  anticommutativity,  and  the  fact 
that  c  x  c  =  0  for  any  vector  c  e  R3.  (Why?)  Thus, 

a  x  b  =  —a2b\k  +  «3i>ij  +  a\b2k  —  a^b2i  —  a\bi]  +  #2^31 

=  (a2b3  -  a3b2)\  +  (a3bi  -  +  {axb2  -  a2bx)k.  (2) 

EXAMPLE  2   Formula  (2)  gives 

(i  +  3j  -  2k)  x  (2i  +  2k)  =  (3  •  2  -  (-2)  •  0)i  +  (-2  •  2  -  1  •  2)j 

+  (1  •  0  -  3  •  2)k 
=  6i  -  6j  -  6k.  ♦ 


Chapter  1  |  Vectors 


Formula  (2)  is  more  complicated  than  the  corresponding  formulas  for  all  the 
other  arithmetic  operations  of  vectors  that  we've  seen.  Moreover,  it  is  a  rather  dif- 
ficult formula  to  remember.  Fortunately,  there  is  a  more  elegant  way  to  understand 
formula  (2).  We  explore  this  reformulation  next. 


Matrices  and  Determinants:  A  First  Introduction  — 
A  matrix  is  a  rectangular  array  of  numbers.  Examples  of  matrices  are 


1  2  3 
4    5  6 


"l 

3 

2 

7 

and 

0 

0 

1  0    0  0" 

0  10  0 

0  0  10 

0  0    0  1 


If  a  matrix  has  m  rows  and  n  columns,  we  call  it  "m  x  n"  (read  "m  by  w"). 
Thus,  the  three  matrices  just  mentioned  are,  respectively,  2  x  3,3  x  2,  and  4  x  4. 
To  some  extent,  matrices  behave  algebraically  like  vectors.  We  discuss  some 
elementary  matrix  algebra  in  §  1 .6.  For  now,  we  are  mainly  interested  in  the  notion 
of  a  determinant,  which  is  a  real  number  associated  toann  x  «  (square)  matrix. 
(There  is  no  such  thing  as  the  determinant  of  a  nonsquare  matrix.)  In  fact,  for  the 
purposes  of  understanding  the  cross  product,  we  need  only  study  2x2  and  3x3 
determinants. 


DEFINITION  4.2    Let  A  be  a  2  x  2  or  3  x  3  matrix.  Then  the  determinant 

of  A,  denoted  det  A  or  |  A|,  is  the  real  number  computed  from  the  individual 
entries  of  A  as  follows: 


•  2  x  2  case 
If  A  = 

•  3  x  3  case 


,  then  |A|  = 


a 

b  c 

d 

e  f 

,  then 

J 

h  i 

a  b 

c 

\A 

d  e 

f 

g  h 

i 

=  ad  —  be. 


=  aei  +bfg  +  cdh  —  ceg  —  afh  —  bdi 


e  f 
h  i 


d  f 
g  i 


+  c 


d  e 
g  h 


in  terms  of  2  x  2  determinants. 


Perhaps  the  easiest  way  to  remember  and  compute  2x2  and  3x3  determi- 
nants (but  not  higher-order  determinants)  is  by  means  of  a  "diagonal  approach." 
We  write  (or  imagine)  diagonal  lines  running  through  the  matrix  entries.  The 
determinant  is  the  sum  of  the  products  of  the  entries  that  lie  on  the  same  diagonal, 


1.4  |  The  Cross  Product 


where  negative  signs  are  inserted  in  front  of  the  products  arising  from  diagonals 
going  from  lower  left  to  upper  right: 


•2x2  case 


A  = 


and 


\A\  =  ad  —  be. 

3x3  case  (we  need  to  repeat  the  first  two  columns  for  the  method  to  work) 

a  b  c 

d  e  f 

_  g  h  i 

Write 


A  = 


Then 


\A\  =  aei  +  bfg  +  cdh  —  gee  —  hfa  —  idb. 


Important  Warning  This  mnemonic  device  does  not  generalize  beyond 
3x3  determinants. 

We  now  state  the  connection  between  determinants  and  cross  products. 


Key  Fact.  If  a  =  a\\  +  d2]  +  a^k  and  b  =  b\i  +  bi\  +  ^k,  then 

j  + 


a  x  b 


a2  a3 
b2  h 


bx  b3 


b\  b2 


k  = 


i    j  k 

a,\  a2  «3 
b\    b2  b} 


(3) 


The  determinants  arise  from  nothing  more  than  rewriting  formula  (2).  Note 
that  the  3  x  3  determinant  in  formula  (3)  needs  to  be  interpreted  by  using  the  2  x  2 
determinants  that  appear  in  formula  (3).  (The  3x3  determinant  is  sometimes 
referred  to  as  a  "symbolic  determinant.") 


EXAMPLE  3 

(3i  +  2j-k)x(i-j  +  k) 


2  -1 

3  -1 

j  + 

3  2 

1  1 

i  — 

1  1 

1  -1 

=  i  -  4j  -  5k 


32       Chapter  1  |  Vectors 


We  may  also  calculate  the  3  x  3  determinant  as 


=  2i  -  j  -  3k  -  2k  -  i  -  3j  =  i  -  4j  -  5k. 


Areas  and  Volumes   

Cross  products  are  used  readily  to  calculate  areas  and  volumes  of  certain  objects. 
We  illustrate  the  ideas  involved  with  the  next  two  examples. 

EXAMPLE  4  Let's  use  vectors  to  calculate  the  area  of  the  triangle  whose  ver- 
tices are  A(3,  1),  B(2,  —1),  and  C(0,  2)  as  shown  in  Figure  1.55. 


B 

Figure  1 .55 

Triangle  ABC  in 

Example  4. 

Figure  1 .56  Any  triangle  may  be 
considered  to  be  half  of  a 
parallelogram. 

The  trick  is  to  recognize  that  any  triangle  can  be  thought  of  as  half  of  a 
parallelogram  (see  Figure  1.56)  and  that  the  area  of  a  parallelogram  is  obtained 
from  a  cross  product.  In  other  words,  A  B  x  A  C  is  a  vector  whose  length  measures 
the  area  of  the  parallelogram  determined  by       and  AC,  and  so 

Area  of  AABC  =  UlA^  x  ACjl. 


To  use  the  cross  product,  we  must  consider  Tb  and  AC  to  be  vectors  in  R3 . 
is  straightforward:  We  simply  take  the  k-components  to  be  zero.  Thus, 


This 


AB  = 


2j  = 


2j  -  Ok, 


and 


Therefore, 


AC  =  — 3i  +  j  =  -3i  +  j  +  0k. 


Tb 


x  AC  = 


j  k 

-2  0 
1  0 


=  -7k. 


Hence, 

Area  of  AABC  =  i||  -  7k||  =  \.  ♦ 


1.4  |  The  Cross  Product 


There  is  nothing  sacred  about  using  A  as  the  common  vertex.  We  could  just 
as  easily  have  used  B  or  C,  as  shown  in  Figure  1.57.  Then 

Area  of  AABC  =  \\\b\  x  B~C\\  =  ~||(i-l-2j)  x  (-2i  +  3j)||  =  ±||7k||  =  \. 

EXAMPLE  5  Find  a  formula  for  the  volume  of  the  parallelepiped  determined 
by  the  vectors  a,  b,  and  c.  (See  Figure  1.58.) 


a  x  b 

B 

Figure  1.57  The  area  of  AABC 
is  7/2. 


|c||  |cos  6\ 


a 

Figure  1 .58  The  parallelepiped  determined  by  a,  b, 
and  c. 

As  explained  in  §1.3,  the  volume  of  a  parallelepiped  is  equal  to  the  product 
of  the  area  of  the  base  and  the  height.  In  Figure  1.58,  the  base  is  the  parallelogram 
determined  by  a  and  b.  Hence,  its  area  is  ||a  x  b||.  The  vector  a  x  b  is  perpendi- 
cular to  this  parallelogram;  the  height  of  the  parallelepiped  is  ||  c  ||  |  cos  6 1 ,  where  9 
is  the  angle  between  a  x  b  and  c.  (The  absolute  value  is  needed  in  case  9  >  tt/2.) 
Therefore, 


Volume  of  parallelepiped  =  (area  of  base)(height) 

=  ||a  x  b||  ||c||  |cos0| 
=  |(axb)-c|. 


(The  appearance  of  the  cos  9  term  should  alert  you  to  the  fact  that  dot  products 
are  lurking  somewhere.) 

For  example,  the  parallelepiped  determined  by  the  vectors 

a  =  i  +  5  j ,    b  =  — 4i  +  2j,    and    c  =  i  +  j  +  6k 
has  volume  equal  to 

|((i  +  5j)x(-4i  +  2j)).(i  +  j  +  6k)|  =  |22k-(i  +  j  +  6k)| 

=  |22(6)| 

=  132.  ♦ 

The  real  number  (a  x  b)  •  c  appearing  in  Example  5  is  known  as  the  triple 
scalar  product  of  the  vectors  a,  b,  and  c.  Since  |  (a  x  b)  •  c|  represents  the  volume 
of  the  parallelepiped  determined  by  a,  b,  and  c,  it  follows  immediately  that 


|(a  x  b)-c|  =  |(b  x  c)-a|  =  |(c  x  a)-b|. 


34       Chapter  1  |  Vectors 


Figure  1 .59  Turning  a  bolt  with  a 
wrench.  The  torque  on  the  bolt  is 
the  vector  r  x  F. 


Figure  1.60  A 

potato  spinning 
about  an  axis. 


Figure  1 .61  The 

angular  velocity 
vector  a>. 


In  fact,  if  you  are  careful  with  the  right-hand  rule,  you  can  convince  yourself  that 
the  absolute  value  signs  are  not  needed;  that  is, 

(a  x  b)  •  c  =  (b  x  c)  •  a  =  (c  x  a)  •  b.  (4) 

This  is  a  nice  example  of  how  the  geometric  significance  of  a  quantity  can  provide 
an  extremely  brief  proof  of  an  algebraic  property  the  quantity  must  satisfy.  (Try 
proving  it  by  writing  out  the  expressions  in  terms  of  components  to  appreciate 
the  value  of  geometric  insight.) 

We  leave  it  to  you  to  check  the  following  beautiful  (and  convenient)  formula 
for  calculating  triple  scalar  products: 


(a  x  b)  •  c  ■ 


ai 


(22  ^3 
C2  C3 


where  a  =  a\\  +  ay  +  «3k,  b  =  b\\  +      +  bj,k,  and  c  =  c\i  +  C2]  +  c^k. 

Torque   

Suppose  you  use  a  wrench  to  turn  a  bolt.  What  happens  is  the  following:  You 
apply  some  force  to  the  end  of  the  wrench  handle  farthest  from  the  bolt  and  that 
causes  the  bolt  to  move  in  a  direction  perpendicular  to  the  plane  determined  by 
the  handle  and  the  direction  of  your  force  (assuming  such  a  plane  exists).  To 
measure  exactly  how  much  the  bolt  moves,  we  need  the  notion  of  torque  (or 
twisting  force). 

In  particular,  letting  F  denote  the  force  you  apply  to  the  wrench,  we  have 

Amount  of  torque  =  (length  of  wrench)(component  of  F  _L  wrench). 

Let  r  be  the  vector  from  the  center  of  the  bolt  head  to  the  end  of  the  wrench 
handle.  Then 


where  9  is  the  ang 
torque  is  ||r  x  F||, 
as  the  direction  in 
bolt).  Hence,  it  is 
torque  vector  T  is 
Note  that  if  F 
fact  that  if  you  try 


Amount  of  torque  =  ||r||  ||F||  sin#, 

;le  between  r  and  F.  (See  Figure  1.59.)  That  is,  the  amount  of 
and  it  is  easy  to  check  that  the  direction  of  r  x  F  is  the  same 
which  the  bolt  moves  (assuming  a  right-handed  thread  on  the 
quite  natural  to  define  the  torque  vector  T  to  be  r  x  F.  The 
a  concise  way  to  capture  the  physics  of  this  situation, 
is  parallel  to  r,  then  T  =  0.  This  corresponds  correctly  to  the 
to  push  or  pull  the  wrench,  the  bolt  does  not  turn. 


Rotation  of  a  Rigid  Body  

Spin  an  object  (a  rigid  body)  about  an  axis  as  shown  in  Figure  1.60.  What  is  the 
relation  between  the  (linear)  velocity  of  a  point  of  the  object  and  the  rotational 
velocity?  Vectors  provide  a  good  answer. 

First  we  need  to  define  a  vector  co,  the  angular  velocity  vector  of  the  rotation. 
This  vector  points  along  the  axis  of  rotation,  and  its  direction  is  determined  by 
the  right-hand  rule.  The  magnitude  of  (o  is  the  angular  speed  (measured  in  radians 
per  unit  time)  at  which  the  object  spins.  Assume  that  the  angular  speed  is  constant 
in  this  discussion.  Next,  fix  a  point  O  (the  origin)  on  the  axis  of  rotation,  and  let 
r(r)  =  OP~  be  the  position  vector  of  a  point  P  of  the  body,  measured  as  a  function 
of  time,  as  in  Figure  1.61.  The  velocity  v  of  P  is  defined  by 

Ar 

v  =  hm  — , 

Ar^O  At 


1.4  |  The  Cross  Product  35 


t(0\ 

Figure  1 .62  A  spinning  rigid 
body. 


where  Ar  =  r(t  +  At)  —  r(t)  (i.e.,  the  vector  change  in  position  between  times  t 
and  t  +  At).  Our  goal  is  to  relate  v  and  co. 

As  the  body  rotates,  the  point  P  (at  the  tip  of  the  vector  r)  moves  in  a  circle 
whose  plane  is  perpendicular  to  co.  (See  Figure  1 .62,  which  depicts  the  motion 
of  such  a  point  of  the  body.)  The  radius  of  this  circle  is  ||r(?)||  sin#,  where  0  is 
the  angle  between  co  and  r.  Both  ||r(/)||  and  6  must  be  constant  for  this  rotation. 
(The  direction  of  r(t)  may  change  with  t,  however.)  If  At  «  0,  then  ||Ar||  is 
approximately  the  length  of  the  circular  arc  swept  by  P  between  /  and  t  +  At. 
That  is, 

||  Ar  ||  «s  (radius  of  circle)(angle  swept  through  by  P) 
=  (||r||  sin0)(A0) 


from  the  preceding  remarks.  Thus, 


Ar 


At 


sinO 


A<p 


Now,  let  At  ->  0.  Then  Ar/Ar  ->  v  and  Acp/At 
angular  velocity  vector  co,  and  we  have 


co    r  sine 


(o  x  r 


\co\\  by  definition  of  the 


(5) 


Figure  1 .63  A  carousel  wheel. 


It's  not  difficult  to  see  intuitively  that  v  must  be  perpendicular  to  both  co  and  r. 
A  moment's  thought  about  the  right-hand  rule  should  enable  you  to  establish  the 
vector  equation 


v  =  co  x  r. 


(6) 


If  we  apply  formula  (5)  to  a  bicycle  wheel,  it  tells  us  that  the  speed  of  a  point 
on  the  edge  of  the  wheel  is  equal  to  the  product  of  the  radius  of  the  wheel  and 
the  angular  speed  (9  is  n/2  in  this  case).  Hence,  if  the  rate  of  rotation  is  kept 
constant,  a  point  on  the  rim  of  a  large  wheel  goes  faster  than  a  point  on  the  rim 
of  a  small  one.  In  the  case  of  a  carousel  wheel,  this  result  tells  you  to  sit  on  an 
outside  horse  if  you  want  a  more  exciting  ride.  (See  Figure  1.63.) 


Summary  of  Products  Involving  Vectors 

Following  is  a  collection  of  some  basic  information  concerning  scalar  multipli- 
cation of  vectors,  the  dot  product,  and  the  cross  product: 


Scalar  Multiplication:  ka 

Result  is  a  vector  in  the  direction  of  a. 
Magnitude  is  ||&a||  =  \k\  ||a||. 
Zero  if  k  =  0  or  a  =  0. 
Commutative:  ka  =  ak. 
Associative:  k(la)  =  (kl)a. 

Distributive:  £(a  +  b)  =  ka  +  kb;  (k  +  l)a  =  ka  +  la. 


Chapter  1  |  Vectors 


Dot  Product:  a  •  b 

Result  is  a  scalar 

Magnitude  is  a  •  b  =  Hall  1  hi  cos  (9* 

G  is  the  an)?1e  between  a  and  b 

rVTapnitnde  is  maximized  ifa  II  Yt 

Zero  if  a  _L  b,  a  =  0,  or  b  =  0. 

Commutative:  a  •  b  =  b  •  a. 

Associativity  is  irrelevant,  since  (a  ■ 

b)  •  c  doesn't  make  sense. 

Distributive:  a-(b  +  c)  =  a-  b  +  a 

•  c. 

Ifa  =  b,  then  a- a  =  ||a||2. 

Cross  Product:  a  x  b 

Result  is  a  vector  perpendicular  to  both  a  and  b. 

Magnitude  is  ||a  x  b||  =  ||a||  ||b||  sin#;  0  is  the  angle  between  a  and  b. 

Magnitude  is  maximized  if  a  _L  b. 

Zero  ifa  ||  b,  a  =  0,  orb  =  0. 

Anticommutative:  axb  =  — bxa. 

Not  associative:  In  general,  a  x  (b  x  c)  ^  (a  x  b)  x  c. 

Distributive:  ax(b  +  c)  =  axb  +  axc  and 
(a  +  b)xc  =  axc  +  bxc. 

Ifa  _Lb,  then  ||axb||  =  ||a||  ||b||. 


Addendum:  Proofs  of  Cross  Product  Properties   

Proof  of  Property  1  To  prove  the  anticommutativity  property,  we  use  the  right- 
hand  rule.  Since 

||axb||  =  ||a||  ||b||  sin0, 

we  obviously  have  that  ||a  x  b||  =  ||b  x  a||.  Therefore,  we  need  only  understand 
the  relation  between  the  direction  of  a  x  b  and  that  of  b  x  a.  To  determine  the 
direction  of  a  x  b,  imagine  curling  the  fingers  of  your  right  hand  from  a  toward  b. 
Then  your  thumb  points  in  the  direction  of  a  x  b.  If  instead  you  curl  your  fingers 
from  b  toward  a,  then  your  thumb  will  point  in  the  opposite  direction.  This  is  the 
direction  of  b  x  a,  so  we  conclude  that  a  x  b  =  — b  x  a.  (See  Figure  1.64.)  ■ 


ax  b 


Figure  1 .64  The  right-hand  rule  shows  why  a  x  b  =  — b  x  a. 


1.4  |  The  Cross  Product  37 
Proof  of  Property  2  First,  note  the  following  general  fact: 


PROPOSITION  4.3  Let  a  and  b  be  vectors  in  R3 .  If  a  •  x  =  b  •  x  for  all  vectors  x 
in  R3,  then  a  =  b. 


To  establish  Proposition  4.3,  write  a  as  aii  +  ci2]  +  03k  and  b  as  b\i  +  + 
b^k  and  set  x  in  turn  equal  to  i,  j,  and  k.  Proposition  4.3  is  valid  for  vectors  in  R2 
as  well  as  R3 . 

To  prove  the  distributive  law  for  cross  products  (property  2),  we  show  that, 
for  any  x  e  R3 , 

(a  x  (b  +  c))  •  x  =  (a  x  b  +  a  x  c)  •  x. 

By  Proposition  4.3,  property  2  follows. 
From  the  equations  in  (4), 

(a  x  (b  +  c))  •  x  =  (x  x  a)  •  (b  +  c) 

=  (x  x  a)  •  b  +  (x  x  a)  •  c, 

from  the  distributive  law  for  dot  products, 

=  (a  x  b)  •  x  +  (a  x  c)  •  x 
=  (a  x  b  +  a  x  c)  •  x, 

again  using  (4)  and  the  distributive  law  for  dot  products.  ■ 

Proof  of  Property  3  Property  3  follows  from  properties  1  and  2.  We  leave  the 
details  as  an  exercise.  ■ 

Proof  of  Property  4  The  second  equality  in  property  4  follows  from  the  first 
equality  and  property  1 : 

k(a  x  b)  =  —k(b  x  a)       by  property  1 

=  —(kb)  x  a       by  the  first  equality  of  property  4 
=  a  x  (kb)         by  property  1 . 

Hence,  we  need  only  prove  the  first  equality. 

If  either  a  or  b  is  the  zero  vector  or  if  a  is  parallel  to  b,  then  the  first  equality 
clearly  holds.  Otherwise,  we  divide  into  three  cases:  (1)  k  =  0,  (2)  k  >  0,  and 
(3)  k  <  0.  If  k  =  0,  then  both  ka  and  k(a  x  b)  are  equal  to  the  zero  vector  and 
the  desired  result  holds.  If  k  >  0,  the  direction  of  (ka)  x  b  is  the  same  as  a  x  b, 
which  is  also  the  same  as  k(a  x  b).  Moreover,  the  angle  between  ka  and  b  is  the 
same  as  between  a  and  b.  Calling  this  angle  0,  we  check  that 

||(*a)xb||  =  ||*a||  ||b||  sine 

=  &||a||  ||b||  sin#  by  part  1  of  Proposition  3.4 

=  k  ||  a  x  b  ||  by  Definition  4. 1 

=  \\k(a  x  b)||  by  part  1  of  Proposition  3.4. 

We  conclude  (ka)  x  b  =  k(a  x  b)  in  this  case. 


38       Chapter  1  |  Vectors 


a 


fca,  k  <  0 


fca,  k  >  0 


Figure  1 .65  If  the  angle  between 
a  and  b  is  0,  then  the  angle 
between  £a  and  b  is  either  0  (if 
k  >  0)  or?r  -0  (if£  <  0). 


If  k  <  0,  then  the  direction  of  (ka)  x  b  is  the  same  as  that  of  (—a)  x  b,  which 
is  seen  to  be  the  same  as  that  of  —(a  x  b)  and  thus  the  same  as  that  of  k(a  x  b). 
The  angle  between  ka  and  b  is  therefore  n  —  9,  where  9  is  the  angle  between  a 
and  b.  (See  Figure  1.65.)  Thus, 

||(*a)  x  b||  =  ||&a||  ||b||  sin(7r  -  9)  =  \k\  ||a||  ||b||  sin  9  =  ||*(a  x  b)||. 

So,  again,  it  follows  that  (ka)  x  b  =  k(a  x  b).  ■ 


1.4  Exercises 


Evaluate  the  determinants  in  Exercises  1—4. 


1. 


3. 


2. 


4. 


In  Exercises  5—7,  calculate  the  indicated  cross  products,  using 
both  formulas  (2)  and  (3). 

5.  (1,  3, -2)  x  (-1,5,  7) 

6.  (3i-2j  +  k)x(i  +  j  +  k) 

7.  (i  +  j)x(-3i  +  2j) 

8.  Prove  property  3  of  cross  products,  using  properties  1 
and  2. 

9.  If  a  x  b  =  3i  -  7j  -  2k,  what  is  (a  +  b)  x  (a  -  b)? 

1 0.  Calculate  the  area  of  the  parallelogram  having  vertices 
(1,1),  (3,  2),  (1,3),  and  (-1,2). 

1 1 .  Calculate  the  area  of  the  parallelogram  having  vertices 
(1,2,  3),  (4,  -2,  1),  (-3,  1,  0),  and  (0,  -3,  -2). 

1 2.  Find  a  unit  vector  that  is  perpendicular  to  both  2i  + 
j  —  3k  and  i  +  k. 

13.  If  (a  x  b)  •  c  =  0,  what  can  you  say  about  the  geomet- 
ric relation  between  a,  b,  and  c? 

Compute  the  area  of  the  triangles  described  in  Exercises 
14-17. 

1 4.  The  triangle  determined  by  the  vectors  a  =  i  +  j  and 
b  =  2i-j 

15.  The  triangle  determined  by  the  vectors  a  =  i  —  2j  + 
6k  and  b  =  4i  +  3j  -  k 

16.  The  triangle  having  vertices  (1,1),  (—1,2),  and 
(-2,-1) 

17.  The  triangle  having  vertices  (1,0,  1),  (0,2,3),  and 
(-1,5,-2) 


1 8.  Find  the  volume  of  the  parallelepiped  determined  by 
a  =  3i  -  j,  b  =  -2i  +  k,  and  c  =  i  -  2j  +  4k. 

1 9.  What  is  the  volume  of  the  parallelepiped  with  vertices 
(3,0,-1),  (4,2,-1),  (-1,1,0),  (3,1,5),  (0,3,0), 
(4,3,5),  (-1,2,  6),  and  (0,4,6)? 


20.  Verify  that  (a  x  b)  •  c 


a  i 


a2  a-} 
b2  b3 
C2  c3 


21 .  Show  that  (a  x  b)  •  c  =  a  •  (b  x  c)  using  Exercise  20. 

22.  Use    geometry    to    show    that    |(axb)-c|  = 
|b-(axc)|. 

23.  (a)  Show  that  the  area  of  the  triangle  with  vertices 

P\(xi,y\),  P2(x2,  yi),  and  P3(x3,  y3)  is  given  by 
the  absolute  value  of  the  expression 


1     1  1 

x\    x2  JC3 

yi  yi  yi 


(b)  Use  part  (a)  to  find  the  area  of  the  triangle  with 
vertices  (1,2),  (2,  3),  and  (-4,  -4). 

24.  Suppose  that  a,  b,  and  c  are  noncoplanar  vectors  in  R3 , 
so  that  they  determine  a  tetrahedron  as  in  Figure  1 .66. 


Figure  1.66  The  tetrahedron  of 
Exercise  24. 

Give  a  formula  for  the  surface  area  of  the  tetrahedron 
in  terms  of  a,  b,  and  c.  (Note:  More  than  one  formula 
is  possible.) 


1.4  |  Exercises  39 


25.  Suppose  that  you  are  given  nonzero  vectors  a,  b,  and  c 
in  R3 .  Use  dot  and  cross  products  to  give  expressions  for 
vectors  satisfying  the  following  geometric  descriptions: 

(a)  A  vector  orthogonal  to  a  and  b 

(b)  A  vector  of  length  2  orthogonal  to  a  and  b 

(c)  The  vector  projection  of  b  onto  a 

(d)  A  vector  with  the  length  of  b  and  the  direction  of  a 

(e)  A  vector  orthogonal  to  a  and  b  x  c 

(f)  A  vector  in  the  plane  determined  by  a  and  b  and 
perpendicular  to  c. 

26.  Suppose  a,  b,  c,  and  d  are  vectors  in  R3 .  Indicate  which 
of  the  following  expressions  are  vectors,  which  are 
scalars,  and  which  are  nonsense  (i.e.,  neither  a  vector 
nor  a  scalar). 


(a)  (a  x  b)  x  c 

(c)  (a  •  b)  x  (c  •  d) 

(e)  (a  •  b)  x  (c  x  d) 

(g)  (a  x  b)  •  (c  x  d) 


(b)  (a-b)-c 
(d)  (a  x  b)  •  c 
(f)  ax[(b-c)d] 
(h)  (a  •  b)c  -  (a  x  b) 


Exercises  27—32  concern  several  identities  for  vectors  a,  b,  c, 
and  d  in  R3.  Each  of  them  can  be  verified  by  hand  by  writing 
the  vectors  in  terms  of  their  components  and  by  using  formula 
(2)  for  the  cross  product  and  Definition  3.1  for  the  dot  product. 
However,  this  is  quite  tedious  to  do.  Instead,  use  a  computer 
algebra  system  to  define  the  vectors  a,  b,  C,  and  d  in  general 
and  to  verify  the  identities. 

^27.  (a  x  b)  x  c  =  (a  •  c)b  -  (b  •  c)a 

^  28.  a  •  (b  x  c)  =  b  •  (c  x  a)  =  c  •  (a  x  b) 

=  -a  •  (c  x  b)  =  -c  •  (b  x  a) 
=  -b  •  (a  x  c) 

^  29.  (a  x  b)  •  (c  x  d)  =  (a  •  c)(b  •  d)  -  (a  •  d)(b  •  c) 
a  •  c   a  •  d 
b-c  b-d 

^  30.  (a  x  b)  x  c  +  (b  x  c)  x  a  +  (c  x  a)  x  b  =  0  (this  is 
known  as  the  Jacobi  identity). 

^  31.  (a  x  b)  x  (c  x  d)  =  [a  •  (c  x  d)]b  -  [b  •  (c  x  d)]a 
^  32.  (a  x  b)  •  (b  x  c)  x  (c  x  a)  =  [a  •  (b  x  c)]2 

33.  Establish  the  identity 

(a  x  b)  •  (c  x  d)  =  (a  •  c)(b  •  d)  -  (a  •  d)(b  •  c) 

of  Exercise  29  without  resorting  to  a  computer  algebra 
system  by  using  the  results  of  Exercises  27  and  28. 

34.  Egbert  applies  a  20  lb  force  at  the  edge  of  a  4  ft 
wide  door  that  is  half-open  in  order  to  close  it.  (See 
Figure  1.67.)  Assume  that  the  direction  of  force  is  per- 
pendicular to  the  plane  of  the  doorway.  What  is  the 
torque  about  the  hinge  on  the  door? 

35.  Gertrude  is  changing  a  flat  tire  with  a  tire  iron.  The  tire 
iron  is  positioned  on  one  of  the  bolts  of  the  wheel  so 


Figure  1.67  Figure  for  Exercise  34. 


40  lb 


Figure  1.68  The  configuration  for 
Exercise  35. 

that  it  makes  an  angle  of  30°  with  the  horizontal.  (See 
Figure  1.68.)  Gertrude  exerts  40  lb  of  force  straight 
down  to  turn  the  bolt. 

(a)  If  the  length  of  the  arm  of  the  wrench  is  1  ft,  how 
much  torque  does  Gertrude  impart  to  the  bolt? 

(b)  What  if  she  has  a  second  tire  iron  whose  length  is 
18  in? 

36.  Egbert  is  trying  to  open  a  jar  of  grape  jelly.  The  ra- 
dius of  the  lid  of  the  jar  is  2  in.  If  Egbert  imparts  15  lb 
of  force  tangent  to  the  edge  of  the  lid  to  open  the  jar, 
how  many  ft-lb,  and  in  what  direction,  is  the  resulting 
torque? 

37.  A  50  lb  child  is  sitting  on  one  end  of  a  seesaw,  3  ft 
from  the  center  fulcrum.  (See  Figure  1 .69.)  When  she  is 


1.5  ft 


Figure  1.69  The  seesaw  of  Exercise  37. 


40       Chapter  1  |  Vectors 


1 .5  ft  above  the  horizontal  position,  what  is  the  amount 
of  torque  she  exerts  on  the  seesaw? 

38.  For  this  problem,  note  that  the  radius  of  the  earth  is 
approximately  3960  miles. 

(a)  Suppose  that  you  are  standing  at  45°  north  latitude. 
Given  that  the  earth  spins  about  its  axis,  how  fast 
are  you  moving? 

(b)  How  fast  would  you  be  traveling  if,  instead,  you 
were  standing  at  a  point  on  the  equator? 

39.  Archie,  the  cockroach,  and  Annie,  the  ant,  are  on  an 
LP  record.  Archie  is  at  the  edge  of  the  record  (ap- 
proximately 6  in  from  the  center)  and  Annie  is  2  in 
closer  to  the  center  of  the  record.  How  much  faster  is 
Archie  traveling  than  Annie?  (Note:  A  record  playing 
on  a  turntable  spins  at  a  rate  of  33  j  revolutions  per 
minute.) 

40.  A  top  is  spinning  with  a  constant  angular  speed  of  12 
radians/sec.  Suppose  that  the  top  spins  about  its  axis 


of  symmetry  and  we  orient  things  so  that  this  axis  is 
the  z-axis  and  the  top  spins  counterclockwise  about  it. 

(a)  If,  at  a  certain  instant,  a  point  P  in  the  top  has 
coordinates  (2,-1,  3),  what  is  the  velocity  of  the 
point  at  that  instant? 

(b)  What  are  the  (approximate)  coordinates  of  P  one 
second  later? 

41.  There  is  a  difficulty  involved  with  our  definition  of 
the  angular  velocity  vector  go,  namely,  that  we  cannot 
properly  consider  this  vector  to  be  "free"  in  the  sense 
of  being  able  to  parallel  translate  it  at  will.  Consider 
the  rotations  of  a  rigid  body  about  each  of  two  parallel 
axes.  Then  the  corresponding  angular  velocity  vectors 
go1  and  <w2  are  parallel.  Explain,  perhaps  with  a  fig- 
ure, that  even  if  G0l  and  G02  are  equal  as  "free  vectors," 
the  corresponding  rotational  motions  that  result  must 
be  different.  (Therefore,  when  considering  more  than 
one  angular  velocity,  we  should  always  assume  that  the 
axes  of  rotation  pass  through  a  common  point.) 


Figure  1 .70  The  plane  in  R3 
through  the  point  Pq  and 
perpendicular  to  the  vector  n. 


1 .5  Equations  for  Planes;  Distance  Problems 

In  this  section,  we  use  vectors  to  derive  analytic  descriptions  of  planes  in  R3.  We 
also  show  how  to  solve  a  variety  of  distance  problems  involving  "fiat  objects" 
(i.e.,  points,  lines,  and  planes). 

Coordinate  Equations  of  Planes   

A  plane  IT  in  R3  is  determined  uniquely  by  the  following  geometric  information: 
a  particular  point  Pq(xo,  vo,  zo)  in  the  plane  and  a  particular  vector  n  =  Ai  + 
B\  +  Ck  that  is  normal  (perpendicular)  to  the  plane.  In  other  words,  n  is  the 

set  of  all  points  P(x,  y,  z)  in  space  such  that  PqP  is  perpendicular  to  n.  (See 
Figure  1.70.)  This  means  that  n  is  defined  by  the  vector  equation 


n-  Kf>  =  0. 


(1) 


Since  PqP  =  (x  —  xo)i  +  (y  —  yo)j  +  (z  —  zo)k,  equation  (1)  may  be  rewritten 
as 

(Ai  +  fij  +  Ck)  •  ((x  -  x0)i  +  (y  -  y0)j  +  (z  -  z0)k)  =  0 

or 


A(x  -  x0)  +  B(y  -  y0)  +  C(z  -  z0)  =  0. 


(2) 


This  is  equivalent  to 

Ax  +  By  +  Cz  =  D, 
where  D  =  Axq  +  Byo  +  Czo- 


1.5  |  Equations  for  Planes;  Distance  Problems  41 


EXAMPLE  1  Theplanethroughthepoint(3,2, 1)  with  normal  vector  2i  —  j  +  4k 
has  equation 

(2i  -  j  +  4k)  •  ((x  -  3)i  +  (y  -  2)j  +  (z  -  l)k)  =  0 
<=►  2(x  -  3)  -  (y  -  2)  +  4(z  -  1)  =  0 

2x  -  y  +  4z  =  8.  # 

Not  only  does  a  plane  in  R3  have  an  equation  of  the  form  given  by  equation 
(2),  but,  conversely,  any  equation  of  this  form  must  describe  a  plane.  Moreover, 
it  is  easy  to  read  off  the  components  of  a  vector  normal  to  the  plane  from  such  an 
equation:  They  are  just  the  coefficients  of  x,  y,  and  z. 

EXAMPLE  2  Given  the  plane  with  equation  Ix  +  2y  —  3z  =  1,  find  a  normal 
vector  to  the  plane  and  identify  three  points  that  lie  on  that  plane. 

A  possible  normal  vector  is  n  =  7i  +  2j  —  3k.  However,  any  nonzero  scalar 
multiple  of  n  will  do  just  as  well.  Algebraically,  the  effect  of  using  a  scalar  multiple 
of  n  as  normal  is  to  multiply  equation  (2)  by  such  a  scalar. 

Finding  three  points  in  the  plane  is  not  difficult.  First,  let  y  =  z  =  0  in  the 
defining  equation  and  solve  for  x: 

Ix  +  2-0-3-0=1     <^=>     7x  =  l  x=f 

Thus  (j,  0,  0)  is  a  point  on  the  plane.  Next,  let  x  =  z  =  0  and  solve  for  y: 

7-0  + 2^-3-0=1  y=\. 

So  (0,  j,  0)  is  another  point  on  the  plane.  Finally,  let  x  =  y  =  0  and  solve  for  z. 
You  should  find  that  (0,0,—  j)  lies  on  the  plane.  ♦ 

EXAMPLE  3  Put  coordinate  axes  on  R3  so  that  the  z-axis  points  vertically. 
Then  a  plane  in  R3  is  vertical  if  its  normal  vector  n  is  horizontal  (i.e.,  if  n  is 
parallel  to  the  xy-plane).  This  means  that  n  has  no  k-component,  so  n  can  be 
written  in  the  form  Ai  +  Bj.  It  follows  from  equation  (2)  that  a  vertical  plane  has 
an  equation  of  the  form 

A(x  -  x0)  +  B(y  -  yo)  =  0. 
Hence,  a  nonvertical  plane  has  an  equation  of  the  form 

A(x  -  x0)  +  B(y  -  y0)  +  C(z  -  z0)  =  0, 
where  C  /  0.  ♦ 

EXAMPLE  4  From  high  school  geometry,  you  may  recall  that  a  plane  is 
determined  by  three  (noncollinear)  points.  Let's  find  an  equation  of  the  plane 
that  contains  the  points  P0(l,  2,  0),  Pi(3,  1,  2),  and  P2(0,  1,  1). 

There  are  two  ways  to  solve  this  problem.  The  first  approach  is  algebraic 
and  rather  uninspired.  From  the  aforementioned  remarks,  any  plane  must  have 
an  equation  of  the  form  Ax  +  By  +  Cz  =  D  for  suitable  constants  A,  B,  C,  and 
D.  Thus,  we  need  only  to  substitute  the  coordinates  of  Po,  Pi,  and  P2  into  this 
equation  and  solve  for  A,  B,  C,  and  D.  We  have  that 

•  substitution  of  Po  gives  A  +  2B  =  D; 

•  substitution  of  Pi  gives  3  A  +  B  +  2C  =  D;  and 

•  substitution  of  P2  gives  B  +  C  =  D. 


42       Chapter  1  |  Vectors 


Figure  1.71  The  plane 
determined  by  the  points  Pq,  P\, 
and  P2  in  Example  4. 


Hence,  we  must  solve  a  system  of  three  equations  in  four  unknowns: 

A  +  2B  =  D 

3 A  +   B  +  2C  =  D  . 
B  +   C  =  D 


(3) 


In  general,  such  a  system  has  either  no  solution  or  else  infinitely  many  solutions. 
We  must  be  in  the  latter  case,  since  we  know  that  the  three  points  Pq,  P\,  and  P2 
lie  on  some  plane  (i.e.,  that  some  set  of  constants  A,  B,  C,  and  D  must  exist). 
Furthermore,  the  existence  of  infinitely  many  solutions  corresponds  to  the  fact 
that  any  particular  equation  for  a  plane  may  be  multiplied  by  a  nonzero  constant 
without  altering  the  plane  defined.  In  other  words,  we  can  choose  a  value  for  one 
of  A,  B,  C,  or  D,  and  then  the  other  values  will  be  determined.  So  let's  multiply 
the  first  equation  given  in  (3)  by  3,  and  subtract  it  from  the  second  equation.  We 
obtain 


A+    2B  =  D 

-5B  +2C  =  -2D  . 
B  +   C  =  D 


(4) 


Now,  multiply  the  third  equation  in  (4)  by  5  and  add  it  to  the  second: 

A  +  2B  =  D 

1C  =  3D  .  (5) 
B  +   C  =  D 

Multiply  the  third  equation  appearing  in  (5)  by  2  and  subtract  it  from  the  first: 

A         -2C  =  -D 

1C=  3D  .  (6) 
B  +      C  =  D 

By  adding  appropriate  multiples  of  the  second  equation  to  both  the  first  and  third 
equations  of  (6),  we  find  that 


1C  = 


3D 


(7) 


B 


=  %D 


Thus,  if  in  (7)  we  take  D  =  —1  (for  example),  then  A  =  1,  B  =  —4,  C  =  —3, 
and  the  equation  of  the  desired  plane  is 

x  -  4y  -  3z  =  -7. 

The  second  method  of  solution  is  cleaner  and  more  geometric.  The  idea  is 
to  make  use  of  equation  (1).  Therefore,  we  need  to  know  the  coordinates  of  a 
particular  point  on  the  plane  (no  problem — we  are  given  three  such  points)  and 

a  vector  n  normal  to  the  plane.  The  vectors  P0P1  and  P0P2  both  lie  in  the  plane. 
(See  Figure  1.71.)  In  particular,  the  normal  vector  n  must  be  perpendicular  to 
them  both.  Consequently,  the  cross  product  provides  just  what  we  need.  That  is, 
we  may  take 


n  =  P0Pi  x  P0P2  =  (2i  -  j  +  2k)  x  (-i  -  j  +  k) 


=  i  -  4j  -  3k. 


1.5  |  Equations  for  Planes;  Distance  Problems  43 


2y  +  z  =  4 


Figure  1 .72  The  line  of 

intersection  of  the  planes 

x  —  2y  +  z  =  4  and 

2x  +  y  +  3z  =  —7  in  Example  5. 


If  we  take  Po(l,  2,  0)  to  be  the  particular  point  in  equation  (1),  we  find  that  the 
equation  we  desire  is 

(i-4j-3k)-((x-  l)i  +  (v-2)j  +  zk)  =  0 

or 

(x  -  l)  -  My  -  2)  -  3z  =  0. 
This  is  the  same  equation  as  the  one  given  by  the  first  method.  ♦ 

EXAMPLE  5    Consider  the  two  planes  having  equations  x  —  2 y  +  z  =  4  and 

2x  +  y  +  3z  =  —  7.  We  determine  a  set  of  parametric  equations  for  their  line  of 
intersection.  (See  Figure  1.72.)  We  use  Proposition  2.1.  Thus,  we  need  to  find  a 
point  on  the  line  and  a  vector  parallel  to  the  line.  To  find  the  point  on  the  line, 
we  note  that  the  coordinates  (x,  y,  z)  of  any  such  point  must  satisfy  the  system  of 
simultaneous  equations  given  by  the  two  planes 


x-2y  +  z=  4 
2x  +  y  +  3z  =  -7 


(8) 


From  the  equations  given  in  (8),  it  is  not  too  difficult  to  produce  a  single 
solution  (x,  y,  z).  For  example,  if  we  let  z  =  0  in  (8),  we  obtain  the  simpler 
system 


(9) 
-2, 


The  solution  to  the  system  of  equations  (9)  is  readily  calculated  to  be  x  = 
y  =  —3.  Thus,  (—2,  —3,  0)  are  the  coordinates  of  a  point  on  the  line. 

To  find  a  vector  parallel  to  the  line  of  intersection,  note  that  such  a  vector 
must  be  perpendicular  to  the  two  normal  vectors  to  the  planes.  The  normal  vectors 
to  the  planes  are  i  —  2j  +  k  and  2i  +  j  +  3k.  Therefore,  a  vector  parallel  to  the 
line  of  intersection  is  given  by 

(i  -  2j  +  k)  x  (2i  +  j  +  3k)  =  -7i  -  j  +  5k. 

Hence,  Proposition  2.1  implies  that  a  vector  parametric  equation  for  the  line  is 

r(r)  =  (-2i  -  3j)  +  f(-7i  -  j  +  5k), 

and  a  standard  set  of  parametric  equations  is 

x  =  -7f  -  2 
y  =  -t-3  . 

z  =  5t  ♦ 


Parametric  Equations  of  Planes   

Another  way  to  describe  a  plane  in  R3  is  by  a  set  of  parametric  equations.  First, 
suppose  that  a  =  (ct\,  a-i,  a^)  and  b  =  (b\,  b2,  b^)  are  two  nonzero,  nonparallel 
vectors  in  R3.  Then  a  and  b  determine  a  plane  in  R3  that  passes  through  the 
origin.  (See  Figure  1.73.)  To  find  the  coordinates  of  a  point  P(x,  y,  z)  in  this 
plane,  draw  a  parallelogram  whose  sides  are  parallel  to  a  and  b  and  that  has  two 
opposite  vertices  at  the  origin  and  at  P,  as  shown  in  Figure  1 .74.  Then  there  must 
exist  scalars  s  and  t  so  that  the  position  vector  of  P  is  ssi  +  tb.  The  plane 


44       Chapter  1  |  Vectors 


z 


Figure  1 .73  The  plane  through 
the  origin  determined  by  the 
vectors  a  and  b. 


z 


y 


Figure  1 .74  For  the  point  P  in 
the  plane  shown,  OP  =  .?a  +  fb 
for  appropriate  scalars  s  and  t. 


Figure  1 .75  The  plane  passing 
through  Pq(c\ ,  C2,  C3)  and  parallel 
to  a  and  b. 


may  be  described  as 

{x  e  R3  |  x  =  sa  +  fb;  s,  t  e  R} . 

Now,  suppose  that  we  seek  to  describe  a  general  plane  n  (i.e.,  one  that  does 
not  necessarily  pass  through  the  origin).  Let 

c  =  (ci,  c2,  c3)  =  OPq 

denote  the  position  vector  of  a  particular  point  P0  in  n  and  let  a  and  b  be  two 
(nonzero,  nonparallel)  vectors  that  determine  the  plane  through  the  origin  parallel 
to  n.  By  parallel  translating  a  and  b  so  that  their  tails  are  at  the  head  of  c  (as  in 
Figure  1.75),  we  adapt  the  preceding  discussion  to  see  that  the  position  vector  of 
any  point  P(x,  y,  z)  in  n  may  be  described  as 

=  sa  +  tb  +  c. 
To  summarize,  we  have  shown  the  following: 

PROPOSITION  5.1  A  vector  parametric  equation  for  the  plane  n  containing 
the  point  Pq(c\,  c2,  c?,)  (whose  position  vector  is  OPq  =  c)  and  parallel  to  the 
nonzero,  nonparallel  vectors  a  and  b  is 

x(s,  t)  =  sa+tb  +  c.  (10) 


By  taking  components  in  formula  (10),  we  readily  obtain  a  set  of  parametric 
equations  for  n : 


x  =  sa\  +  tb\  +  C\ 

y  =  sa2  +  tb2  +  c2  ■  (11) 
Z  =  scii  +  tb-i  +  c3 


Compare  formula  (10)  with  that  of  equation  (1)  in  Proposition  2.1.  We  need 
to  use  two  parameters  s  and  t  to  describe  a  plane  (instead  of  a  single  parameter 
t  that  appears  in  the  vector  parametric  equation  for  a  line)  because  a  plane  is  a 
two-dimensional  object. 


1.5  |  Equations  for  Planes;  Distance  Problems  45 


EXAMPLE  6   We  find  a  set  of  parametric  equations  for  the  plane  that  passes 
through  the  point  (1,0,-1)  and  is  parallel  to  the  vectors  3i  —  k  and  2i  +  5j  +  2k. 
From  formula  (10),  any  point  on  the  plane  is  specified  by 


x(s,  t)  =  s(3i  -  k)  +  r(2i  +  5j  +  2k)  +  (i 
=  (3s  +  2t  +  l)i  +  5t\  +  (2/  -  s  - 
The  individual  parametric  equation  may  be  read  off  as 

x  =  3s  +  2t  +  1 


-k) 
l)k. 


y  =  5t 
z  =  2t 


1 


Figure  1.76  A  general 
configuration  for  finding  the 
distance  between  a  point  and  a 
line,  using  vector  projections. 


Figure  1 .77  Another  general 
configuration  for  finding  the 
distance  between  a  point  and  a 
line. 


Distance  Problems   

Cross  products  and  vector  projections  provide  convenient  ways  to  understand  a 
range  of  distance  problems  involving  lines  and  planes:  Several  examples  follow. 
What  is  important  about  these  examples  are  the  vector  techniques  for  solving 
geometric  problems  that  they  exhibit,  not  the  general  formulas  that  may  be  derived 
from  them. 

EXAMPLE  7  (Distance  between  a  point  and  a  line)  We  find  the  distance 
between  the  point  P0(2,  1 ,  3)  and  the  line  1(0  =  t(- 1 ,  1 ,  -2)  +  (2,  3 ,  -2)  in  two 
ways. 

Method  1.  From  the  vector  parametric  equations  for  the  given  line,  we  read 
off  a  point  B  on  the  line — namely,  (2,  3,  —2) — and  a  vector  a  parallel  to  the 
line — namely,  a  =  (—  1,  1,-2).  Using  Figure  1.76,  the  length  of  the  vector 
BPq  —  projaSPo  provides  the  desired  distance  between  P0  and  the  line.  Thus, 
we  calculate  that 

B~?q  =  {2,  1,3) -(2,  3,-2) 
=  (0,  -2,  5); 


projaBP0  = 


a- BP, 


o 


a  •  a 


(-1,  1,-2). (0,-2,  5) 


(-1,1,-2) 


"  V("l,l,-2)-(-l,l,-2X 
=  (2,  -2,  4). 

The  desired  distance  is 

||BP0-projafiP0||  =  11(0,  -2,  5) -(2, -2,  4)||  =  ||(-2,  0, 

Method  2.  In  this  case,  we  use  a  little  trigonometry.  If  0  denotes  the  angle 
between  the  vectors  a  and  BPq  as  in  Figure  1.77,  then 

D 

sin  6*  = 


=  V~5. 


\\BP0\ 


where  D  denotes  the  distance  between  Pq  and  the  line.  Hence, 

|| a ||  || Mil  sine  HaxBPol 


D=  \\BP0\\  sin<9  = 


a 


46       Chapter  1  |  Vectors 


Therefore,  we  calculate 

a  x  B~fi)  = 

so  that  the  distance  sought  is 
D  = 


i     j  k 

-1  1  -2 
0-2  5 


|i  +  5j  +  2k|| 


-i  +  j-2k||  76 
which  agrees  with  the  answer  obtained  by  Method  1 . 


=  i  +  5j  +  2k, 


Figure  1 .78  The  general 
configuration  for  finding  the 
distance  D  between  two  parallel 
planes. 


/CJL — c 

Bo 


Figure  1.79  Configuration  for 
determining  the  distance  between 
two  skew  lines  in  Example  9. 


EXAMPLE  8  (Distance  between  parallel  planes)  The  planes 
111:  2x -2y+z  =  5    and    n2:  2x  -  2y  +  z  =  20 

are  parallel.  (Why?)  We  see  how  to  compute  the  distance  between  them. 

Using  Figure  1.78  as  a  guide,  we  see  that  the  desired  distance  D  is  given  by 
||projn/,iiD2||,  where  Pi  is  a  point  on  Oi,  P2  is  a  point  on  n2,  and  n  is  a  vector 
normal  to  both  planes. 

First,  the  vector  n  that  is  normal  to  both  planes  may  be  read  directly  from 
the  equation  for  either  n  1  or  Tl2  as  n  =  2i  —  2j  +  k.  It  is  not  hard  to  find  a  point 
Pi  on  TI 1 :  the  point  Pi(0,  0,  5)  will  do.  Similarly,  take  P2(0,  0,  20)  for  a  point  on 
n2.  Then 


P,P2  =  (0,0,  15), 


and  calculate 


projnP1P2  = 


n-PiP2 


n  •  n 


n  = 


(2,-2,  1).  (0,0,  15) 


(2. 


=  -¥(2, 


-2,1) -(2, 

v-,  "2,  1) 
=  -§(2,-2,1). 
Hence,  the  distance  D  that  we  seek  is 
D 


-2,  1) 


(2,-2,1) 


llprojnPi/>2 


\V9 


EXAMPLE  9  (Distance  between  two  skew  lines)  Find  the  distance  between 
the  two  skew  lines 

li(0  =  *(2.  1,3)  +  (0,  5,-1)    and   l2(r)  =  f(l,  -1,  0)  +  (-1, 2,  0). 

(Two  lines  in  R3  are  said  to  be  skew  if  they  are  neither  intersecting  nor  parallel. 
It  follows  that  the  lines  must  lie  in  parallel  planes  and  that  the  distance  between 
the  lines  is  equal  to  the  distance  between  the  planes.) 

To  solve  this  problem,  we  need  to  find  ||projn5i  B2 1| ,  the  length  of  the  projec- 
tion of  the  vector  between  a  point  on  each  line  onto  a  vector  n  that  is  perpendicular 
to  both  lines,  hence,  also  perpendicular  to  the  parallel  planes  that  contain  the  lines. 
(See  Figure  1.79.) 

From  the  vector  parametric  equations  for  the  lines,  we  read  that  the  point 
Bi(0,  5,  —1)  is  on  the  first  line  and  52(—  1,  2,  0)  is  on  the  second.  Hence, 


B,B2  =  (-1,  2,  0)  -  (0,  5,  -1)  =  (-1,  -3,  1). 


1.5  |  Exercises  47 


For  a  vector  n  that  is  perpendicular  to  both  lines,  we  may  use  n  =  ai  x  a2,  where 
ai  =  (2,  1 ,  3)  is  a  vector  parallel  to  the  first  line  and  a2  =  ( 1 ,  —  1 ,  0)  is  parallel  to 
the  second.  (We  may  read  these  vectors  from  the  parametric  equations.)  Thus, 


n  =  ai  x  a2 


and  so, 


projn£i£2 


n-  BiB2 


n  •  n 


J 


=  3i  +  3j  -  3k, 


(-1,-3,  l)-(3,3,-3) 
(3,3,  -3) -(3,  3,  -3) 


(3,3,-3) 


(3,3, 


=  -f(l,l, 


-3) 
!)• 


The  desired  distance  is  ||projnJSi  Z?2  II  =  |V3. 


1.5  Exercises 


1 .  Calculate  an  equation  for  the  plane  containing  the  point 
(3,-1,2)  and  perpendicular  to  i  —  j  +  2k. 

2.  Find  an  equation  for  the  plane  containing  the  point 
(9,  5,  —1)  and  perpendicular  to  i  —  2k. 

3.  Find  an  equation  for  the  plane  containing  the  points 
(3,-1,2),  (2,  0,5),  and  (1, -2,4). 

4.  Find  an  equation  for  the  plane  containing  the  points 
(A,  0,  0),  (0,  B,  0),  and  (0,  0,  C).  Assume  that  at  least 
two  of  A,  B,  and  C  are  nonzero. 

5.  Give  an  equation  for  the  plane  that  is  parallel  to  the 
plane  5x  —  4y  +  z  =  1  and  that  passes  through  the 
point  (2,  —1,  —2). 

6.  Give  an  equation  for  the  plane  parallel  to  the  plane  2x  — 
3y  +  z  =  5  that  passes  through  the  point  (—  1 ,  1,2). 

7.  Find  an  equation  for  the  plane  parallel  to  the  plane  x  — 
y  +  Iz  =  10  that  passes  through  the  point  (—2,  0,  1). 

8.  Give  an  equation  for  the  plane  parallel  to  the  plane 
2x  +  2y  +  z  =  5  and  that  contains  the  line  with  para- 
metric equations  x  =  2  —  t,  y  =  2t  +  1,  z  =  3  —  2t. 

9.  Explain  why  there  is  no  plane  parallel  to  the  plane 
5x  —  3  y  +  2z  =  10  that  contains  the  line  with  para- 
metric equations  x  =  t  +  4,  y  =  3t  —  2,  z  =  5  —  2t . 

1 0.  Find  an  equation  for  the  plane  that  contains  the  line  x  = 
2t  -  1,  y  =  3t  +  4,  z  =  1  -  t  and  the  point  (2,  5,  0). 

11.  Find  an  equation  for  the  plane  that  is  perpendicular 
to  the  line  x  =  It  -  5,  y  =  7  -  It,  z  =  8  -  t  and  that 
passes  through  the  point  (1,  —1,2). 


12.  Find  an  equation  for  the  plane  that  contains  the  two 
lines  l\.x  =  t  +  2,  y  =  3t  —  5,z  =  5t  +  1  andh'.x  = 

5  -  t,  y  =  3t  -  10,  z  =  9  -  2t. 

1 3.  Give  a  set  of  parametric  equations  for  the  line  of  inter- 
section of  the  planes  x  +  2y  —  3z  =  5  and  5x  +  5y  — 
z  =  1. 

1 4.  Give  a  set  of  parametric  equations  for  the  line  through 
(5,  0,  6)  that  is  perpendicular  to  the  plane  2x  —  3y  + 
5z  =  -1. 

1 5.  Find  a  value  for  A  so  that  the  planes  Sx  —  6y  +  9  Az  = 

6  and  Ax  +  y  +  2z  =  3  are  parallel. 

16.  Find  values  for  A  so  that  the  planes  Ax  —  y  +  z  =  1 
and  3  Ax ;  +  Ay  —  2z  =  5  are  perpendicular. 

Give  a  set  of  parametric  equations  for  each  of  the  planes  de- 
scribed in  Exercises  1 7—22. 

17.  The  plane  that  passes  through  the  point  (—1,  2,  7)  and 
is  parallel  to  the  vectors  2i  —  3j  +  k  and  i  —  5k 

18.  The  plane  that  passes  through  the  point  (2,9,-4) 
and  is  parallel  to  the  vectors  —  8i  +  2j  +  5k  and 
3i  -  4j  -  2k 

19.  The  plane  that  contains  the  lines  h:  x  =  2t  +  5,  y  = 

-3t  -  6,z  =  4f  +  10  and  l2:  x  =  5t  -  1,  y  =  lOt  + 
3,z  =  lt-2 

20.  The  plane  that  passes  through  the  three  points  (0,  2,  1), 
(7, -1,5),  and  (-1,3,0) 

21.  The  plane  that  contains  the  line  /:  x  =  3t  —  5,  y  = 
10  -  3t,  z  =  2t  +  9  and  the  point  (-2,  4,  7) 


48       Chapter  1  |  Vectors 


22.  The  plane  determined  by  the  equation  2x  —  3y  + 
5z  =  30 

23.  Find  a  single  equation  of  the  form  Ax  +  By  +  Cz  =  D 
that  describes  the  plane  given  parametrically  as  x  = 
3s  -  t  +  2,  y  =  4s  +  t,  z  =  s  +  5t  +  3.  (Hint:  Begin 
by  writing  the  parametric  equations  in  vector  form  and 
then  find  a  vector  normal  to  the  plane.) 

24.  Find  the  distance  between  the  point  (1 ,  —2,  3)  and  the 
line  I:  x  =  2t  -  5,  y  =  3  -  t,  z  =  4. 

25.  Find  the  distance  between  the  point  (2,  —1)  and  the 

line/:jc  =  3f  +  7,  y  =  5t  -  3. 

26.  Find  the  distance  between  the  point  (—11,  10,  20)  and 
the  line  /:  x  =  5  -  t ,  y  =  3,  z  =  It  +  8. 

27.  Determine  the  distance  between  the  two  lines  li(f)  = 
f(8, -1,0) +  (-1,3,  5)  and  I2(f)  =  f(0,  3, 1)  + 
(0,  3,  4). 

28.  Compute  the  distance  between  the  two  lines 
h(0  =  (t  ~  7)i  +(5f  +  l)j  +  (3  -2f)k  and  l2(r)  = 
4ri  +  (2-0j  +  (8f+l)k. 

29.  (a)  Find  the  distance  between  the  two  lines  li(f)  = 

f(3, 1,2) +  (4, 0,2)  and  I2(f)  =  t(\,  2,  3)  + 
(2,1,3). 

(b)  What  does  your  answer  in  part  (a)  tell  you  about 
the  relative  positions  of  the  lines? 

30.  (a)  The  lines  h(r)  =  t(l, -1,  5)  +  (2,  0,  -4)  and 

l2(r)  =  ?(l,-l,5)  +  (l,3,-5)  are  parallel.  Ex- 
plain why  the  method  of  Example  9  cannot  be  used 
to  calculate  the  distance  between  the  lines. 

(b)  Find  another  way  to  calculate  the  distance.  (Hint: 
Try  using  some  calculus.) 

31.  Find  the  distance  between  the  two  planes  given  by  the 
equations  x  —  3 y  +  2z  =  1  and  x  —  3y  +  2z  =  8. 

32.  Calculate  the  distance  between  the  two  planes 

5x  -  2y  +  2z  =  12    and     -  10.v  +  4y  -  4z  =  8. 

33.  Show  that  the  distance  d  between  the  two  parallel 
planes  determined  by  the  equations  Ax  +  By  +  Cz  = 


Di  and  Ax  +  By  +  Cz  =  D2  is 
d=  \Di-D2\ 

Va2  +  b2  +  c2' 

34.  Two  planes  are  given  parametrically  by  the  vector 
equations 

Xl(s,  f)  =  (-3,  4,  -9)  +  s(9,  -5,  9)  +  f(3,  -2,  3) 
x2(s,  t)  =  (5,  0,  3)  +  s(-9, 2,  -9)  +  r(-4,  7,  -4). 

(a)  Give  a  convincing  explanation  for  why  these 
planes  are  parallel. 

(b)  Find  the  distance  between  the  planes. 

35.  Write  equations  for  the  planes  that  are  parallel  to 
x  +  3y  —  5z  =  2  and  lie  three  units  from  it. 

36.  Suppose  that  li(r)  =  fa  +  bi  and  l2(f)  =  ra  +  b2  are 
parallel  lines  in  either  R2  or  R3 .  Show  that  the  distance 
D  between  them  is  given  by 

D=  ||ax(b2-bi)|| 
11*11 

(Hint:  Consider  Example  7.) 

37.  Let  n  be  the  plane  in  R3  with  normal  vector  n  that 
passes  through  the  point  A  with  position  vector  a.  If  b 
is  the  position  vector  of  a  point  B  in  R3,  show  that  the 
distance  D  between  B  and  n  is  given  by 

|n-(b-a)| 

D  =  -. 

1 1  n  1 1 

38.  Show  that  the  distance  D  between  parallel  planes  with 
normal  vector  n  is  given  by 

|n-(x2  -xQ| 
1 1  n  1 1 

where  X]  is  the  position  vector  of  a  point  on  one  of  the 
planes,  and  x2  is  the  position  vector  of  a  point  on  the 
other  plane. 

39.  Suppose  that  li  (f)  =  t&i  +  bi  andl2(f)  =  fa2  +  b2  are 
skew  lines  in  R3 .  Use  the  geometric  reasoning  of  Ex- 
ample 9  to  show  that  the  distance  D  between  these  lines 
is  given  by 

_  |(at  x  a2)-(b2  -bQ| 
||  ai  x  a2|| 


1 .6   Some  n-dimensional  Geometry 

Vectors  in  R"   

The  algebraic  idea  of  a  vector  in  R2  or  R3  is  defined  in  §  1 . 1 ,  in  which  we  asked 
you  to  consider  what  would  be  involved  in  generalizing  the  operations  of  vector 
addition,  scalar  multiplication,  etc.,  to  n-dimensional  vectors,  where  n  can  be 
arbitrary.  We  explore  some  of  the  details  of  such  a  generalization  next. 


1.6  |  Some  n-dimensional  Geometry  49 


DEFINITION  6.1  A  vector  in  R"  is  an  ordered  rc-tuple  of  real  numbers.  We 
use  a  =  (til,  a2,  •  •  ■ ,  an)  as  our  standard  notation  for  a  vector  in  R". 


EXAMPLE  1  The  5-tuple  (2,  4,  6,  8,  10)  is  a  vector  in  R5.  The  (n  +  l)-tuple 
(2«,  2rc  —  2,  2«  —  4,  . . . ,  2,  0)  is  a  vector  in  R"+1 ,  where  «  is  arbitrary.  ♦ 

Exactly  as  is  the  case  in  R2  or  R3 ,  we  call  two  vectors  a  =  (a\ ,  a2 , . . . ,  an)  and 

b  =  (pi,  b2  Z?„)  equal  just  incases,  =  bt  for/  =  1,2,...,  n.  Vector  addition 

and  scalar  multiplication  are  defined  in  complete  analogy  with  Definitions  1 .3  and 
1.4:  If  a  =  (a\,  a2, . . . ,  an)  and  b  =  (b\ ,  b2, . . . ,  b„)  are  two  vectors  in  R"  and 
k  e  R  is  any  scalar,  then 

a  +  b  =  (ai  +  b\,  a2  +  b2  an  +  fc„) 

and 

k&  =  (ka\,  ka2, . . . ,  ^an). 

The  properties  of  vector  addition  and  scalar  multiplication  given  in  §1.1  hold 
(with  proofs  that  are  no  different  from  those  in  the  two-  and  three-dimensional 
cases).  Similarly,  the  dot  product  of  two  vectors  in  R"  is  readily  defined: 

a  •  b  =  fli&i  +  a2b2  +  •  •  •  +  a„bn. 

The  dot  product  properties  given  in  §1.3  continue  to  hold  in  n  dimensions;  we 
leave  it  to  you  to  check  that  this  is  so. 

What  we  cannot  do  in  dimensions  larger  than  three  is  to  develop  a  pictorial 
representation  for  vectors  as  arrows.  Nonetheless,  the  power  of  our  algebra  and 
analogy  does  allow  us  to  define  a  number  of  geometric  ideas.  We  define  the  length 
of  a  vector  in  a  e  R"  by  using  the  dot  product: 

||a||  =  Va^a. 

The  distance  between  two  vectors  a  and  b  in  R"  is 

Distance  between  a  and  b  =  ||a  —  b||. 

We  can  even  define  the  angle  between  two  nonzero  vectors  by  using  a  generalized 
version  of  equation  (4)  of  §  1 .3 : 


Here  a,  be  R"  and  9  is  taken  so  that  0  <  6  <  it.  (Note:  At  this  point  in  our 
discussion,  it  is  not  clear  that  we  have 

a-b 

-1  <   <  1, 

"  Ila||  ||b||  " 

which  is  a  necessary  condition  if  our  definition  of  the  angle  9  is  to  make  sense. 
Fortunately,  the  Cauchy-Schwarz  inequality — formula  (1)  that  follows — takes 
care  of  this  issue.)  Thus,  even  though  we  are  not  able  to  draw  pictures  of  vectors 
in  R",  we  can  nonetheless  talk  about  what  it  means  to  say  that  two  vectors  are 
perpendicular  or  parallel,  or  how  far  apart  two  vectors  may  be.  (Be  careful  about 
this  business.  We  are  defining  notions  of  length,  distance,  and  angle  entirely  in 


Chapter  1  |  Vectors 


terms  of  the  dot  product.  Results  like  Theorem  3.3  have  no  meaning  in  R",  since 
the  ideas  of  angles  between  vectors  and  dot  products  are  not  independent.) 

There  is  no  simple  generalization  of  the  cross  product.  However,  see  Exer- 
cises 39^12  at  the  end  of  this  section  for  the  best  we  can  do  by  way  of  analogy. 

We  can  create  a  standard  basis  of  vectors  in  R"  that  generalize  the  i,  j, 
k-basis  in  R3 .  Let 

d  =(1,0,  0.....0), 
e2=(0,  1,0,...,  0), 


e„  =  (0,  0, . . . ,  0,  1). 
Then  it  is  not  difficult  to  see  (check  for  yourself)  that 

a  =  (ai,  a2, . . . ,  a„)  =  a\&\  +  a2e2  H  h  a„e„ 

Here  are  two  famous  (and  often  handy)  inequalities: 


PROOF  If  n  =  2  or  3,  this  result  is  virtually  immediate  in  view  of  Theorem  3.3. 
However,  in  dimensions  larger  than  three,  we  do  not  have  independent  notions  of 
inner  products  and  angles,  so  a  different  proof  is  required. 

First  note  that  the  inequality  holds  if  either  a  or  b  is  0.  So  assume  that  a  and 
b  are  nonzero.  Then  we  may  define  the  projection  of  b  onto  a  just  as  in  §1.3: 

/a-b\ 
projab  =  I  —  I  a  =  ka. 

Here  k  is,  of  course,  the  scalar  a  •  b/a  •  a.  Let  c  =  b  —  ka  (so  that  b  =  ka  +  c). 
Then  we  have  a  •  c  =  0,  since 


a  •  c 


a  •  a 


=  a 

(b- 

-  jfca) 

=  a 

b- 

ka  •  a 
/a-b 

=  a 

b- 

\a  •  a 

=  a 

b- 

a-b 

=  0. 

We  leave  it  to  you  to  check  that  the  "Pythagorean  theorem"  holds,  namely,  that 
the  following  equation  is  true: 


|b||2  =  £2||a||2  +  ||c|| 


Multiply  this  equation  by  ||  a 

2 


\2  - 


a  •  a.  We  obtain 


a 


b||2  = 


|a||2£2||a||2  +  ||a||2||c||2 


=  a 


a-b 


a  •  a 


|a||2  +  ||a||2||c||2 


1.6  |  Some  n-dimensional  Geometry 


=  (»-»)(^)  (a.a)+||a||2||c||2 

=  (a.b)2  +  ||a||2||c||2. 
Now,  the  quantity  ||a||2||c||2  is  nonnegative.  Hence, 

||a||2||b||2>(a.b)2. 
Taking  square  roots  in  this  last  inequality  yields  the  result  desired.  ■ 

The  geometric  motivation  for  this  proof  of  the  Cauchy-Schwarz  inequality 
comes  from  Figure  1 .80. 1 


PROOF  Strategic  use  of  the  Cauchy-Schwarz  inequality  yields 
||a  +  b||2  =  (a  +  b)-(a  +  b) 

=  a-  a  +  2a-b  +  b-  b 

<  a-a  +  2||a||  ||b|| +b-b  by(l) 
=  ||a||2  +  2||a||  ||b||  +  ||b||2 
=  (l|a||  +  ||b||)2. 

Thus,  the  result  desired  holds  by  taking  square  roots,  since  the  quantities  on  both 
sides  of  the  inequality  are  nonnegative.  ■ 

In  two  or  three  dimensions  the  triangle  inequality  has  the  following  obvious 
proof  from  which  the  inequality  gets  its  name:  Since  ||a||,  ||b||,  and  ||a  +  b||  can 
be  viewed  as  the  lengths  of  the  sides  of  a  triangle,  inequality  (2)  says  nothing 
more  than  that  the  sum  of  the  lengths  of  two  sides  of  a  triangle  must  be  at  least 
as  large  as  the  length  of  the  third  side,  as  demonstrated  by  Figure  1.81. 

Matrices   

We  had  a  brief  glance  at  matrices  and  determinants  in  §  1 .4  in  connection  with  the 
computation  of  cross  products.  Now  it's  time  for  another  look. 

A  matrix  is  defined  in  §  1 .4  as  a  rectangular  array  of  numbers.  To  extend  our 
discussion,  we  need  a  good  notation  for  matrices  and  their  individual  entries.  We 
used  the  uppercase  Latin  alphabet  to  denote  entire  matrices  and  will  continue  to 
do  so.  We  shall  also  adopt  the  standard  convention  and  use  the  lowercase  Latin 
alphabet  and  two  sets  of  indices  (one  set  for  rows,  the  other  for  columns)  to 
identify  matrix  entries.  Thus,  the  general  m  x  n  matrix  can  be  written  as 


A  = 


a\\  an  ■•■  din 
C?21    an    ■  ■  ■  ain 

-  "ml     @m2     '  '  '  @mn 


=  (shorthand)  (fly). 


1  See  J.  W.  Cannon,  Amer.  Math.  Monthly  96  (1989),  no.  7,  630-631. 


Chapter  1  |  Vectors 


The  first  index  always  will  represent  the  row  position  and  the  second  index,  the 
column  position. 

Vectors  in  R"  can  also  be  thought  of  as  matrices.  We  shall  have  occasion  to 
write  the  vector  a  =  (a\,  a%, . . . ,  an)  either  as  a  row  vector  (a  1  x  n  matrix), 

a  =  [  a\   ai   ■  ■  ■   an  ] , 

or,  more  typically,  as  a  column  vector  (an  nxl  matrix), 

a\ 

a2 

a  = 


1-    "71  — I 


We  did  not  use  double  indices  since  there  is  only  a  single  row  or  column  present. 
It  will  be  clear  from  context  (or  else  indicated  explicitly)  in  which  form  a  vector 
a  will  be  viewed.  An  m  x  n  matrix  A  can  be  thought  of  as  a  "vector  of  vectors" 
in  two  ways:  (1)  as  m  row  vectors  in  R", 

[  an  an  ■  ■  ■  a\n  ] 
[  <221    a22    ■  ■  ■    a2n  ] 


A 


[  am\  am2 
or  (2)  as  n  column  vectors  in  R"! , 


nn  1 


an 

an 

a\n 

a2\ 

a2n 

-  am\  - 

-  am2  _ 

-  amn  _ 

A  = 


We  now  define  the  basic  matrix  operations.  Matrix  addition  and  scalar  mul- 
tiplication are  really  no  different  from  the  corresponding  operations  on  vectors 
(and  moreover,  they  satisfy  essentially  the  same  properties). 


DEFINITION  6.2  (Matrix  Addition)  Let  A  and  B  be  two  m  x  n  matrices. 
Then  their  matrix  sum  A  +  B  is  the  m  x  n  matrix  obtained  by  adding  cor- 
responding entries.  That  is,  the  entry  in  the  ith  row  and  7  th  column  of  A  +  B 
is  atj  +  btj,  where  a,7  and     are  the  ijth  entries  of  A  and  B,  respectively. 


EXAMPLE  2  If 

A  = 

then 


However,  if  B 


and  B 


A  +  B  = 


2 
10 


the  same  dimensions  as  A. 


,  then  A  +  B  is  not  defined,  since  B  does  not  have 


1.6  |  Some  n-dimensional  Geometry 


Properties  of  matrix  addition.  For  all  m  x  n  matrices  A,  B,  and  C  we  have 

1.  A  +  B  =  B  +  A  (commutativity); 

2.  A  +  (B  +  C)  =  (A  +  B)  +  C  (associativity); 

3.  An  m  x  n  matrix  O  (the  zero  matrix)  with  the  property  that  A  +  O  =  A 
for  all  m  x  n  matrices  A. 


DEFINITION  6.3  (Scalar  Multiplication)  If  A  is  an  m  x  n  matrix  and 
k  e  R  is  any  scalar,  then  the  product  kA  of  the  scalar  k  and  the  matrix  A  is 
obtained  by  multiplying  every  entry  in  A  by  k.  That  is,  the  i/th  entry  of  kA 
is  kciij  (where  ay  is  the  iyth  entry  of  A). 


EXAMPLE  3   IfA  = 


1 

4 


,  then  3A  = 


3  6  c 
12    15  If 


Properties  of  scalar  multiplication.  If  A  and  B  are  any  m  x  n  matrices 
and  k  and  /  are  any  scalars,  then 

1.  (k  +  l)A  =  kA  +  lA  (distributivity); 

2.  k(A  +  B)  =  kA  +  kB  (distributivity); 

3.  k(lA)  =  (kl)A  =  l(kA). 


We  leave  it  to  you  to  supply  proofs  of  these  addition  and  scalar  multiplication 
properties  if  you  wish. 

Just  as  defining  products  of  vectors  needed  to  be  "unexpected"  in  order  to  be 
useful,  so  it  is  with  defining  products  of  matrices.  To  a  degree,  matrix  multipli- 
cation is  a  generalization  of  the  dot  product  of  two  vectors. 


DEFINITION  6.4   (Matrix  Multiplication)  Let  A  be  an  m  x  n  matrix 
and  B  an  n  x  p  matrix.  Then  the  matrix  product  A  B  is  the  m  x  p  matrix 
whose  i/th  entry  is  the  dot  product  of  the  j'th  row  of  A  and  the  y'th  column 
of  B  (considered  as  vectors  in  R").  That  is,  the  ijth  entry  of 

au     ayi    ■  ■  ■  a\n 
[an     ai2    ■  ■  ■  ain] 

-bn  ... 

b2\ 

~b,jl 
bij 

. . .  bip 

bip 

.bn\ 

-  bnj  _ 

Kp. 

is 

a 

i\b\j  +  ai2b2j  H  h  a 

'n 

n 

bnj  =  (more  compactly)       gaby  ■ 

k=l 

54       Chapter  1  |  Vectors 


EXAMPLE  4  If 

A  = 


1  2  3 
4     5  6 


and    B  = 


0  1 
7  0 
2  4 


then  the  (2,  1)  entry  of  AB  is  the  dot  product  of  the  second  row  of  A  and  the  first 
column  of  B: 


(2,  1)  entry  =  [  4      5      6  ] 


=  (4)(0)  +  (5)(7)  +  (6)(2)  =  47. 


The  full  product  A  B  is  the  2x2  matrix 


20  13 
47  28 

On  the  other  hand  B  A  is  the  3x3  matrix 


4     5  6 
7    14  21 
18    24  30 


Order  matters  in  matrix  multiplication.  To  multiply  two  matrices  we  must 


have 

Number  of  columns  of  left  matrix  =  number  of  rows  of  right  matrix. 

In  Example  4,  the  products  AB  and  BA  are  matrices  of  different  dimensions; 
hence,  they  could  not  possibly  be  the  same.  A  worse  situation  occurs  when  the 
matrix  product  is  defined  in  one  order  and  not  the  other.  For  example,  if  A  is  2  x  3 
and  £  is  3  x  3,  then  A  B  is  defined  (and  is  a  2  x  3  matrix),  but  B  A  is  not.  However, 
even  if  both  products  AB  and  BA  are  defined  and  of  the  same  dimensions  (as  is 
the  case  if  A  and  B  are  both  n  x  n,  for  example),  it  is  in  general  still  true  that 

AB  +  BA. 

Despite  this  negative  news,  matrix  multiplication  does  behave  well  in  a  number 
of  respects,  as  the  following  results  indicate: 


Properties  of  matrix  multiplication.  Suppose  A,  B,  and  C  are  matrices 
of  appropriate  dimensions  (meaning  that  the  expressions  that  follow  are  all 
defined)  and  that  k  is  a  scalar.  Then 

1.  A(BC)  =  (AB)C; 

2.  k(AB)  =  (kA)B  =  A(kB); 

3.  A(B  +  C)  =  AB  +  AC; 

4.  (A  +  B)C  =  AC  +  BC. 


The  proofs  of  these  properties  involve  little  more  than  Definition  6.4,  although 
the  notation  can  become  somewhat  involved  as  in  the  proof  of  property  1 . 


One  simple  operation  on  matrices  that  has  no  analogue  in  the  real  number 
system  is  the  transpose.  The  transpose  of  an  m  x  n  matrix  A  is  the  n  x  m  matrix 


1.6  |  Some  n-dimensional  Geometry 


AT  obtained  by  writing  the  rows  of  A  as  columns.  For  example,  if 

then    AT  = 


A  = 


1 

4 


More  abstractly,  the  ijth  entry  of  AT  is  a,-,-,  the  jith  entry  of  A. 

The  transpose  operation  turns  row  vectors  into  column  vectors  and  vice  versa 
We  also  have  the  following  results: 

(ATf 


A, 

(A  B)1  =  BTAT, 


for  any  matrix  A.  (3) 
where  A  is  m  x  n  and  B  is  n  x  p.  (4) 


The  transpose  will  largely  function  as  a  notational  convenience  for  us.  For 
example,  consider  a,  b  £  R"  to  be  column  vectors.  Then  the  dot  product  a  •  b  can 
be  written  in  matrix  form  as 


■  b  =  a\b\  +  a2b2  +  •  •  •  +  a„bn  =  [  a\ 


C'2 


In  ] 


b2 


=  arb. 


EXAMPLE  5  Matrix  multiplication  is  defined  the  way  it  is  so  that,  roughly 
speaking,  working  with  vectors  or  quantities  involving  several  variables  can  be 
made  to  look  as  much  as  possible  like  working  with  a  single  variable.  This  idea 
will  become  clearer  throughout  the  text,  but  we  can  provide  an  important  example 
now.  A  linear  function  in  a  single  variable  is  a  function  of  the  form  f(x)  =  ax 
where  a  is  a  constant.  The  natural  generalization  of  this  to  higher  dimensions 
is  a  linear  mapping  F:  R"  — >  R",  F(x)  =  Ax,  where  A  is  a  (constant)  m  x  n 
matrix  and  x  £  R".  More  explicitly,  F  is  a  function  that  takes  a  vector  in  R" 
(written  as  a  column  vector)  and  returns  a  vector  in  R"!  (also  written  as  a  column). 
That  is, 


F(x)  =  Ax  = 


an 

an  ■ 

X\ 

Cl2l 

a22  ■ 

_  am\ 

@m2 

&mn  _ 

The  function  F  has  the  properties  that  F(x  +  y)  =  F(x)  +  F(y)  for  all  x,  y  e  R" 
and  F(kx)  =  k¥(x)  for  all  x  e  R",  i  £  R.  These  properties  are  also  satisfied  by 
f(x)  =  ax,  of  course.  Perhaps  more  important,  however,  is  the  fact  that  linear 
mappings  behave  nicely  with  respect  to  composition.  Suppose  F  is  as  just  defined 
and  G:  R"!  — >  Rp  is  another  linear  mapping  defined  by  G(x)  =  Bx,  where  B  is  a 
p  x  m  matrix.  Then  there  is  a  composite  function  GoF:  R"  — >  Rp  defined  by 

G  o  F(x)  =  G(F(x))  =  G(Ax)  =  fi(Ax)  =  (BA)x 

by  the  associativity  property  of  matrix  multiplication.  Note  that  BA  is  defined 
and  is  a  p  x  n  matrix.  Hence,  we  see  that  the  composition  of  two  linear  mappings 
is  again  a  linear  mapping.  Part  of  the  reason  we  defined  matrix  multiplication  the 
way  we  did  is  so  that  this  is  the  case.  ♦ 


EXAMPLE  6  We  saw  that  by  interpreting  equation  (1)  in  §1.2  inn  dimensions, 
we  obtain  parametric  equations  of  a  line  in  R" .  Equation  (2)  of  §  1 .5,  the  equation 


56       Chapter  1  |  Vectors 

for  a  plane  in  R3  through  a  given  point  (xo,  yo,  zo)  with  given  normal  vector 
n  =  Ai  +  B]  +  Ck,  can  also  be  generalized  to  n  dimensions: 

Ax(xx  -  bi)  +  A2(x2  —  b2)  H  h  An(x„  -  bn)  =  0. 

If  we  let  A  =  (Ai,  A2, . . . ,  A„),  b  =  (b\,  &2>  •  •  •  >  b„)  ("constant"  vectors),  and 
x  =  (jci,  x%,  ■  ■  ■ ,  x„)  (a  "variable"  vector),  then  the  aforementioned  equation  can 
be  rewritten  as 

A  •  (x  -  b)  =  0 

or,  considering  A,  b,  and  x  as  n  x  1  matrices,  as 

Ar(x  -  b)  =  0. 

This  is  the  equation  for  a  hyperplane  in  R"  through  the  point  b  with  normal 
vector  A.  The  points  x  that  satisfy  this  equation  fill  out  an  (n  —  l)-dimensional 
subset  of  R".  ♦ 

At  this  point,  it  is  easy  to  think  that  matrix  arithmetic  and  the  vector  geometry 
of  R",  although  elegant,  are  so  abstract  and  formal  as  to  be  of  little  practical  use. 
However,  the  next  example,  from  the  field  of  economics,2  shows  that  this  is  not 
the  case. 

EXAMPLE  7  Suppose  that  we  have  n  commodities.  If  the  price  per  unit  of  the 
ith  commodity  is  pi,  then  the  cost  of  purchasing  x,  (>  0)  units  of  commodity 
i  is  piXj.  If  p  =  (pi, .. . ,  p„)  is  the  price  vector  of  all  the  commodities  and 
x  =  (jti, . . . ,  xn)  is  the  commodity  bundle  vector,  then 

p  •  x  =  pxx\  +  p2x2  H  V  p»x„ 

represents  the  total  cost  of  the  commodity  bundle. 

Now  suppose  that  we  have  an  exchange  economy,  so  that  we  may  buy  and 
sell  items.  If  you  have  an  endowment  vector  w  =  (w\ ,...,«>„),  where  to,-  is  the 
amount  of  commodity  /  that  you  can  sell  (trade),  then,  with  prices  given  by  the 
price  vector  p,  you  can  afford  any  commodity  bundle  x  where 

p  •  x  <  p  •  w. 

We  may  rewrite  this  last  equation  as 

p  •  (x  -  w)  <  0. 

In  other  words,  you  can  afford  any  commodity  bundle  x  in  the  budget  set 
{x  |  p  •  (x  —  w)  <  0}.  The  equation  p  •  (x  —  w)  =  0  defines  a  budget  hyperplane 
passing  through  w  with  normal  vector  p.  ♦ 


Determinants   

We  have  already  defined  determinants  of  2  x  2  and  3x3  matrices.  (See  §1.4.) 
Now  we  define  the  determinant  of  any  n  x  n  (square)  matrix  in  terms  of  determi- 
nantsof(n  —  1)  x  (n  —  1)  matrices.  By  "iterating  the  definition,"  we  can  calculate 
any  determinant. 


2  See  D.  Saari,  "Mathematical  complexity  of  simple  economics,"  Notices  of  the  American  Mathematical 
Society  42  (1995),  no.  2,  222-230. 


1.6  |  Some  n-dimensional  Geometry 


DEFINITION  6.5  Let  A  =  (ay)  be  an  n  x  n  matrix.  The  determinant  of 
A  is  the  real  number  given  by 

l)1+2«i2|A12|  +  ---  +  (-l)1+Bai„|A1„|, 
1)  submatrix  of  A  obtained  by  deleting  the 


\A\={-X)l+lan\An 


where  Ay  is  the  (n  —  1)  x  (« 
rth  row  and  jth  column  of  A. 


EXAMPLE  8   IfA  = 


According  to  Definition  6.5, 


det 


1 

2 

1 

3 

-2 

1 

0 

5 

4 

2 

-1 

0 

3 

-2 

1 

1 

,  then 


+- 

i— 

—i— 

-2 

0 

5 

4 

> 

-1 

0 

3 

> 

1 

1 

2 

0 

5 

4 

-1 

0 

3 

1 

1 

1 

2 

1 

3  " 

-2 

1 

0 

5 

4 

2 

-1 

0 

3 

-2 

1 

1 

1 

0 

5 

(-l)1+1(l)det 

2 

-1 

0 

-2 

1 

1 

-2 

0 

5 

+  (- 

l)1+2(2)det 

4 

-1 

0 

3 

1 

1 

"  -2 

1 

5 

+  (- 

l)1+3(l)det 

4 

2 

0 

3 

-2 

1 

"  -2 

1 

0 

+  (- 

l)1+4(3)det 

4 

2 

-1 

3 

-2 

1 

(l)(l)(-l)  +  (-l)(2)(37)  +  (l)(l)(-78) 
+  (-l)(3)(-7) 

-132.  ♦ 


The  determinant  of  the  submatrix  A/j  of  A  is  called  the  ijth  minor  of  A,  and 
the  quantity  (— 1),+;  |A,7  |  is  called  the  ijth  cofactor.  Definition  6.5  is  known  as 
cofactor  expansion  of  the  determinant  along  the  first  row,  since  det  A  is  written 
as  the  sum  of  the  products  of  each  entry  of  the  first  row  and  the  corresponding 
cofactor  (i.e.,  the  sum  of  the  terms  a\j  times  (— |Ay  |). 

It  is  natural  to  ask  if  one  can  compute  determinants  by  cofactor  expansion 
along  other  rows  or  columns  of  A.  Happily,  the  answer  is  yes  (although  we  shall 
not  prove  this). 


58       Chapter  1  |  Vectors 


Convenient  Fact.  The  determinant  of  A  can  be  computed  by  cofactor  ex- 

pansion along  any  row  or  column.  That  is, 

\A\  =  (-ly+'anl  An  |  +  (-l)i+2ai2\Ai2\  +  ■  ■ 

■  +  (-i)i+nain\Ain\ 

(expansion  along  the  rth  row), 

|A|  =  (-l)!+^ly.|Aly-|  +  (-l)2+Ja2j \A2j\  +  • 

■  +  {-X)n+>anl\Anj\ 

(expansion  along  the  y'th  column). 

EXAMPLE  9   To  compute  the  determinant  of 


-  1 

2 

0 

4 

5  _ 

2 

0 

0 

9 

0 

7 

5 

1 

-1 

0 

0 

2 

0 

0 

2 

3 

1 

0 

0 

0 

expansion  along  the  first  row  involves  more  calculation  than  necessary.  In  partic- 
ular, one  would  need  to  calculate  four  4x4  determinants  on  the  way  to  finding 
the  desired  5x5  determinant.  (To  make  matters  worse,  these  4x4  determinants 
would  in  turn,  need  to  be  expanded  also.)  However,  if  we  expand  along  the  third 
column,  we  find  that 

det  A  =  (- l)1+3(0)  det  A13  +  (- 1)2+3(0)  det  A23  +  (- 1)3+3(1)  det  A33 

+  (- 1)4+3(0)  det  A43  +  (- 1)5+3(0)  det  A53 

=  det  A33 


1 

2 

4 

5 

2 

0 

9 

0 

0 

2 

0 

2 

3 

1 

0 

0 

There  are  several  good  ways  to  evaluate  this  4x4  determinant.  We'll  expand 
about  the  bottom  row: 


12      4  5 

2  0  9  0 
0      2      0  2 

3  10  0 


2 

4  5 

1 

4  5 

(-1)4+1(3) 

0 

9  0 

+  (-l)4+2(l) 

2 

9  0 

2 

0  2 

0 

0  2 

=  (-l)(3)(-54)  +  (l)(l)(2) 
=  164. 


Of  course,  not  all  matrices  contain  well-distributed  zeros  as  in  Example  9, 
so  there  is  by  no  means  always  an  obvious  choice  for  an  expansion  that  avoids 
much  calculation.  Indeed  one  does  not  compute  determinants  of  large  matrices 
by  means  of  cofactor  expansion.  Instead  certain  properties  of  determinants  are 
used  to  make  hand  computations  feasible.  Since  we  shall  rarely  need  to  consider 
determinants  larger  than  3  x  3,  we  leave  such  properties  and  their  significance  to 
the  exercises.  (See,  in  particular,  Exercises  26  and  27.) 


1.6  |  Exercises  59 


1.6  Exercises 


1 .  Rewrite  in  terms  of  the  standard  basis  for  R" : 

(a)  (1,2,3,  ...,n) 

(b)  (1,0,-1,  1,0,-1,...,  1,0, -1)  (Assume  that  n 
is  a  multiple  of  3.) 

In  Exercises  2—4  write  the  given  vectors  without  recourse  to 
standard  basis  notation. 

2.  ei  +  e2  H  he,, 

3.  ei  -  2e2  +  3e3  -  4e4  H  h  (-l)"+1ne„ 

4.  ei  +  e„ 

5.  Calculate  the  following,  where  a  =  (1,  3,  5, . . . , 
In  -  1)  and  b  =  (2,  -4,  6,  ... ,  (-l)"+12w): 


(a)  a  +  b 
(d)  ||a|| 


(b)  a  -  b 
(e)  a-b 


(c)  -3a 


6.  Let  n  be  an  even  number.  Verify  the  triangle  in- 
equality in  R"  for  a  =  (1,  0,  1,  0, ... ,  0)  and  b  = 
(0,  1,  0,  1, ... ,  1). 

7.  Verify  that  the  Cauchy-Schwarz  inequality  holds  for 
the  vectors  a  =  (1 ,  2, . . . ,  n)  and  b  =  ( 1 ,  1 , . . . ,  1). 

8.  If  a  =  (1,-1,7,  3,  2)  and  b  =  (2,  5,  0,  9,  -1),  calcu- 
late the  projection  projab. 

9.  Show,  for  all  vectors  a,  b,  c  e  R",  that 

|| a  —  b||  <  || a  —  c||  +  ||c  —  b||. 

10.  Prove  the  Pythagorean  theorem.  That  is,  if  a,  b,  and 
c  are  vectors  in  R"  such  that  a  +  b  =  c  and  a  •  b  =  0, 
then 

||a||2  +  ||b||2  =  ||c||2. 
Why  is  this  called  the  Pythagorean  theorem? 

11.  Let  a  and  b  be  vectors  in  R".  Show  that  if  ||a  +  b||  = 
||  a  —  b||,  then  a  and  b  are  orthogonal. 

12.  Let  a  and  b  be  vectors  in  R".  Show  that  if  ||a  —  b||  > 
||  a  +  b  || ,  then  the  angle  between  a  and  b  is  obtuse  (i.e., 
more  than  jt/2). 

13.  Describe  "geometrically"  the  set  of  points  in  R5  satis- 
fying the  equation 

2(Xl  -  1)  +  3(x2  +  2)  -  7x3  +  x4  -  4  -  5(x5  +  1)  =  0. 

14.  To  make  some  extra  money,  you  decide  to  print  four 
types  of  silk-screened  T-shirts  that  you  sell  at  various 
prices.  You  have  an  inventory  of  20  shirts  that  you  can 
sell  for  $8  each,  30  shirts  that  you  sell  for  $10  each,  24 
shirts  that  you  sell  for  $  12  each,  and  20  shirts  that  you 
sell  for  $  1 5  each.  A  friend  of  yours  runs  a  side  business 
selling  embroidered  baseball  caps  and  has  an  inventory 


of  30  caps  that  can  be  sold  for  $10  each,  16  caps  that 
can  be  sold  for  $10  each,  20  caps  that  can  be  sold  for 
$12  each,  and  28  caps  that  can  be  sold  for  $15  each. 
You  suggest  swapping  half  your  inventory  of  each  type 
of  T-shirt  for  half  his  inventory  of  each  type  of  baseball 
cap.  Is  your  friend  likely  to  accept  your  offer?  Why  or 
why  not? 

15.  Suppose  that  you  run  a  grain  farm  that  produces  six 
types  of  grain  at  prices  of  $200,  $250,  $300,  $375, 
$450,  $500  per  ton. 

(a)  If  x  =  (xi,  . . . ,  xg)  is  the  commodity  bundle  vec- 
tor (meaning  that  x,-  is  the  number  of  tons  of  grain 
i  to  be  purchased),  express  the  total  cost  of  the 
commodity  bundle  as  a  dot  product  of  two  vectors 
inR6. 

(b)  A  customer  has  a  budget  of  $100,000  to  be  used 
to  purchase  your  grain.  Express  the  set  of  possible 
commodity  bundle  vectors  that  the  customer  can 
afford.  Also  describe  the  relevant  budget  hyper- 
plane  in  R6. 

In  Exercises  16-19,  calculate  the  indicated  matrix  quantities 
where 


C 


i  : 

-2  ( 

1  -1 

2  0 
0  3 


16.  3A-2S 
18.  DB 


,      D  = 

17.  AC 
19.  BTD 


20.  The  n  x  n  identity  matrix,  denoted  I  or  7„,  is  the  ma- 
trix whose  nth  entry  is  1  and  whose  other  entries  are 
all  zero.  That  is, 


0  0 


(a)  Explicitly  write  out  I2,  h,  and  I4. 

(b)  The  reason  /  is  called  the  identity  matrix  is  that 
it  behaves  as  follows:  Let  A  be  any  m  x  n  matrix. 
Then 

i.  AI„  =  A. 

ii.  ImA  =  A. 

Prove  these  results.  (Hint:  What  are  the  yth  entries 
of  the  products  in  (i)  and  (ii)?) 


Chapter  1  |  Vectors 


Evaluate  the  determinants  given  in  Exercises  21—23. 


22. 


-7 

/ 

A 

u 

1 

—  1 

0 

2 

0 

1 

3 

1 

-3 

0 

2 

0 

5 

1 

-2 

o 
0 

A 

A 
U 

0 

1  c 

1 
1 

A 

u 

0 

7 

0 

1 

—  1 

0 

8 

1 

9 

7 

5 

-1 

0 

8 

11 

0 

2 

1 

9 

7 

0 

0 

4 

-3 

5 

0 

0 

0 

2 

1 

0 

0 

0 

0 

-3 

24.  Prove  that  a  matrix  that  has  a  row  or  a  column  con- 
sisting entirely  of  zeros  has  determinant  equal  to 
zero. 

25.  An  upper  triangular  matrix  is  an  n  x  n  matrix  whose 
entries  below  the  main  diagonal  are  all  zero.  (Note: 
The  main  diagonal  is  the  diagonal  going  from  upper 
left  to  lower  right.)  For  example,  the  matrix 


-  1 

2 

-1 

2  " 

0 

3 

4 

3 

0 

0 

5 

6 

_  0 

0 

0 

7  _ 

is  upper  triangular. 

(a)  Give  an  analogous  definition  for  a  lower  triangu- 
lar matrix  and  also  an  example  of  one. 

(b)  Use  cofactor  expansion  to  show  that  the  determi- 
nant of  any  n  x  n  upper  or  lower  triangular  matrix 
A  is  the  product  of  the  entries  on  the  main  diagonal. 
That  is,  det  A  =  awa^i  ■  ■  ■  a„„ . 

26.  Some  properties  of  the  determinant.  Exercises  24 
and  25  show  that  it  is  not  difficult  to  compute  de- 
terminants of  even  large  matrices,  provided  that  the 
matrices  have  a  nice  form.  The  following  operations 
(called  elementary  row  operations)  can  be  used  to 
transform  annxn  matrix  into  one  in  upper  triangular 
form: 

I.  Exchange  rows  i  and  j. 
II.  Multiply  row   by  a  nonzero  scalar. 

III.  Add  a  multiple  of  row  i  to  row  j.  (Row  i  remains 

unchanged.) 
For  example,  one  can  transform  the  matrix 

"  0      2      3  " 

1      7  -2 
_  1      5      9  _ 

into  one  in  upper  triangular  form  in  three  steps: 


Step  1 .  Exchange  rows  1  and  2  (this  puts  a  nonzero 
entry  in  the  upper  left  corner): 


0 

2 

3  " 

1 

7 

-2 

1 

7 

-2 

0 

2 

3 

1 

5 

9 

1 

5 

9 

Step  2.  Add  —1  times  row  1  to  row  3  (this  eliminates 
the  nonzero  entries  below  the  entry  in  the  upper  left 
corner): 


Step  3 


1 

7  -2 

"  1 

7 

-2 

0 

2  3 

0 

2 

3 

1 

5  9 

0 

-2 

11 

Add  row  2  to  row  3: 

1 

7  -2 

"  1 

7 

-2 

0 

2  3 

0 

2 

3 

0 

-2  11 

0 

0 

14 

The  question  is,  how  do  these  operations  affect  the  de- 
terminant? 

(a)  By  means  of  examples,  make  a  conjecture  as  to  the 
effect  of  a  row  operation  of  type  I  on  the  determi- 
nant. (That  is,  if  matrix  B  results  from  matrix  A  by 
performing  a  single  row  operation  of  type  I,  how  are 
det  A  and  det  B  related?)  You  need  not  prove  your 
results  are  correct. 

(b)  Repeat  part  (a)  in  the  case  of  a  row  operation  of 
type  III. 

(c)  Prove  that  if  B  results  from  A  by  multiplying  the 
entries  in  the  ;'th  row  of  A  by  the  scalar  c  (a  type  II 
operation),  then  det  B  =  c  ■  det  A. 

27.  Calculate  the  determinant  of  the  matrix 


"  2 

1 

-2 

7 

8 

1 

0 

1 

-2 

4 

-1 

1 

2 

3 

-5 

0 

2 

3 

1 

7 

_  -3 

2 

-1 

0 

1 

by  using  row  operations  to  transform  A  into  a  matrix 
in  upper  triangular  form  and  by  using  the  results  of 
Exercise  26  to  keep  track  of  how  the  determinant  of  A 
and  the  determinant  of  your  final  matrix  are  related. 

28.  (a)  Is  det(A  +  B)  =  det  A  +  det  5?  Why  or  why  not? 
(b)  Calculate 

1  2  7 

3+2      1-1  5+1 
0-2  0 

and 


1 

2 

7 

1 

2 

7 

3 

1 

5 

+ 

2 

-1 

1 

0 

-2 

0 

0 

-2 

0 

and  compare  your  results. 


1.6  |  Exercises  61 


1 

3 

2  +  3 

0 

4 

-1+5 

-1 

0 

0-2 

1 

3 

2 

1 

3 

3 

0 

4 

-1 

+ 

0 

4 

5 

-1 

0 

0 

-1 

0 

-2 

(c)  Calculate 


and 


and  compare  your  results. 

(d)  Conjecture  and  prove  a  result  about  sums  of  deter- 
minants. (You  may  wish  to  construct  further  exam- 
ples such  as  those  in  parts  (b)  and  (c).) 

29.  It  is  a  fact  that,  if  A  and  B  are  any  n  x  n  matrices,  then 

det(AB)  =  (detA)(detfi). 

Use  this  fact  to  show  that  det(A5)  =  det(BA).  (Recall 
that  AB  /  BA,  in  general.) 

An  n  x  n  matrix  A  is  said  to  be  invertible  (or  nonsingular)  if 

there  is  another  n  x  n  matrix  B  with  the  property  that 

AB  =  BA  =  I„, 

where  In  denotes  the  n  x  n  identity  matrix.  (See  Exercise  20.) 
The  matrix  B  is  called  an  inverse  to  the  matrix  A.  Exercises 
30-38  concern  various  aspects  of  matrices  and  their  inverses. 


30.  (a)  Verify  that 


(b)  Verify  that 


1  0 
1  1 


is  an  inverse  of 


1  0 
-1  1 


1  2  3 

2  5  3 
1    0  8 


is   an   inverse  of 


-40    16  9 

13   -5  -3 

5   -2  -1 

31.  Using  the  definition  of  an  inverse  matrix,  find  an 

2  2  1 

inverse  to     0  1  0 

0  0-1 


32.  Try  to  find  an  inverse  matrix  to 
What  happens? 


2  1 
1  0 
0  -1 


33.  Show  that  if  an  n  x  n  matrix  A  is  invertible,  then  A  can 
have  only  one  inverse  matrix.  Thus,  we  may  write  A-1 
to  denote  the  unique  inverse  of  a  nonsingular  matrix 
A.  (Hint:  Suppose  A  were  to  have  two  inverses  B  and 
C.  Consider  B(AC).) 

34.  Suppose  that  A  and  B  are  n  x  «  invertible  matrices. 
Show  that  the  product  matrix  A  5  is  invertible  by  ver- 
ifying that  its  inverse  (AS)-1  =  B~l  A~l . 


35.  (a)  Show  that  if  A  is  invertible,  then  det  A  /  0.  (In 

fact,  the  converse  is  also  true.) 

(b)  Show     that     if     A     is     invertible,  then 

det(A"1)=  — — . 
v      ;  detA 

36.  (a)  Show  that,  if  ad  —  be  /  0,  then  a  general  2x2 
a  b 


matrix 


1 


c  d 


has  the  matrix 


ad  —  be 


d 

-c 


ad— be 

c 

ad— be 


as  inverse. 

(b)  Use  this  formula  to  find  an  inverse  of 


ad— be 
a 

ad— be 


2  4 
-1  2 


37.  If  A  is  a  3  x  3  matrix  and  det  A  /  0,  then  there  is 
a  (somewhat  complicated)  formula  for  A-1.  In 
particular, 


detA 


|A„| 
-IA12I 
|An| 


-IA21I 
IA22I 
-IA23I 


IA31I 
-IA32I 
IA33I 


where  Ay  denotes  the  submatrix  of  A  obtained  by 
deleting  the  ;th  row  and  7th  column  (see  Defini- 
tion 6.5).  Use  this  formula  to  find  the  inverse  of 


2    1  1 

0  2  4 

1  0  3 


More  generally,  if  A  is  any  n  x  n  matrix  and  det  A^0, 
then 


detA 


where  adj  A  is  the  adjoint  matrix  of  A,  that  is,  the 
matrix  whose  ijth  entry  is  (—\)1+-'\Aji\.  (Note:  The 
formula  for  the  inverse  matrix  using  the  adjoint  is  typi- 
cally more  of  theoretical  than  practical  interest,  as  there 
are  more  efficient  computational  methods  to  determine 
the  inverse,  when  it  exists.) 

38.  Repeat  Exercise  37  with  the  matrix 


2-1  3 
1  2  -2 
3      0  1 


Cross  products  in  R".  Although  it  is  not  possible  to  define  a 
cross  product  of  two  vectors  in  R"  as  we  did  for  two  vectors 
in  R3,  we  can  construct  a  "cross  product"  of  n  —  1  vectors 
in  R"  that  behaves  analogously  to  the  three-dimensional  cross 


62       Chapter  1  [  Vectors 


product.  To  be  specific,  if 

ai  =  (an,  au,       ain),     a2  =  (021,  022,  ■  ■  • ,  a2„), 

=  (fln-u,  in-12,  ■  ■  ■ ,  a„-\„) 

are  n  —  1  vectors  in  R",  we  define  ai  X  a2  X  •  •  •  X  a„_i  to  be 
the  vector  in  R"  given  by  the  symbolic  determinant 


ai  x  a2  x  •  •  •  x  a„_i 


"21 


a„-i  1 


e2 
"12 
"22 


a„-i2 


e„ 

«2» 


(Here  t\ ,  . . . ,  e„  are  £/ze  standard  basis  vectors  for  R" .)  Exer- 
cises 39-42  concern  this  generalized  notion  of  cross  product. 

39.  Calculate  the  following  cross  product  in  R4: 

(1,  2,  -1,  3)  x  (0,  2,  -3,  1)  x  (-5,  1,  6,  0). 

40.  Use  the  results  of  Exercises  26  and  28  to  show  that 

(a)  ai  x  •  •  •  x  a,-  x  •  •  •  x  a^  x  •  •  •  x  a„_i 

=  —  (ai  x  •  •  •  x  a j  x  •  •  •  x  a,-  x  •  •  •  x  a„_i ), 
I  <i  <n  —  1 , 1 <  j  <  n  —  1 

(b)  ai  x  •  •  •  x  kHi  x  •  •  •  x  a„_i 

=  &(ai  x  •  •  •  x  a,-  x  •  •  •  x  a„_i), 
1  <  i  <  n  —  1 . 


(c)  ai  x  •  •  •  x  (a,-  +  b)  x  •  •  •  x  a„_i 

=  ai  x  •  •  •  x  a,-  x  •  •  •  x  a„_i  + 
aix---xbx---x  a„_i, 
1  <i  <n  -  l.allbeR". 

(d)  Show  that  if  b  =  (b\ , . . . ,  b„)  is  any  vector  in  R" , 
then 

b  •  (ai  x  a2  x  •  •  •  x  a„_i) 
is  given  by  the  determinant 


All 
aB-n 


b„ 


41.  Show  that  the  vector  b  =  ai  x  a2  x  •  •  •  x  a„_i  is 
orthogonal  to  ai , . . . ,  a„_i . 

42.  Use  the  generalized  notion  of  cross  products  to 
find  an  equation  of  the  (four-dimensional)  hyper- 
plane  in  R5  through  the  five  points  Po(l,  0,  3,  0,  4), 
^(2,-1,0,0,5),  P2(7,  0,  0,  2,  0),  P3(2,0,3,0,4), 
and  P4(l,~  1,3,0,  4). 


Figure  1 .82  The  Cartesian 
coordinate  system. 


Location  y 


(x,y) 


Location  x 

Figure  1 .83  Locating  a  point  P, 
using  Cartesian  coordinates. 


1 .7   New  Coordinate  Systems 

We  hope  that  you  are  comfortable  with  Cartesian  (rectangular)  coordinates  for  R2 
or  R3 .  The  Cartesian  coordinate  system  will  continue  to  be  of  prime  importance  to 
us,  but  from  time  to  time,  we  will  find  it  advantageous  to  use  different  coordinate 
systems.  In  R2,  polar  coordinates  are  useful  for  describing  figures  with  circular 
symmetry.  In  R3,  there  are  two  particularly  valuable  coordinate  systems  besides 
Cartesian  coordinates:  cylindrical  and  spherical  coordinates.  As  we  shall  see, 
cylindrical  and  spherical  coordinates  are  each  a  way  of  adapting  polar  coordinates 
in  the  plane  for  use  in  three  dimensions. 

Cartesian  and  Polar  Coordinates  on  R2   

You  can  understand  the  Cartesian  (or  rectangular)  coordinates  (x ,  y)  of  a  point 
P  in  R2  in  the  following  way:  Imagine  the  entire  plane  filled  with  horizontal  and 
vertical  lines,  as  in  Figure  1.82.  Then  the  point  P  lies  on  exactly  one  vertical  line 
and  one  horizontal  line.  The  x -coordinate  of  P  is  where  this  vertical  line  intersects 
the  x-axis,  and  the  y-coordinate  is  where  the  horizontal  line  intersects  the  y-axis. 
(See  Figure  1.83.)  (Of  course,  we've  already  assigned  coordinates  along  the  axes 
so  that  the  zero  point  of  each  axis  is  at  the  point  of  intersection  of  the  axes.  We 
also  normally  mark  off  the  same  unit  distance  on  each  axis.)  Note  that,  because 
of  this  geometry,  every  point  in  R2  has  a  uniquely  determined  set  of  Cartesian 
coordinates. 

Polar  coordinates  are  defined  by  considering  different  geometric  informa- 
tion. Now  imagine  the  plane  filled  with  concentric  circles  centered  at  the  origin 
and  rays  emanating  from  the  origin.  Then  every  point  except  the  origin  lies  on 


1.7  |  New  Coordinate  Systems  63 


Figure  1 .86  Locating  the  point 
with  polar  coordinates  (r,  9), 
where  r  <  0. 


0 

r  =  6  cos  0 

0 

6 

jr/6 

3V3 

n/4 

3V2 

7T/3 

3 

n/2 

0 

2jt/3 

-3 

3jt/4 

-3V2 

5?r/6 

-3V3 

7T 

-6 

77T/6 

-3V3 

57T/4 

-3V2 

4tt/3 

-3 

3^/2 

0 

57T/3 

3 

7tt/4 

372 

Figure  1 .84  The  polar  coordinate 
system. 


Figure  1 .85  Locating  a  point  P, 
using  polar  coordinates. 


exactly  one  such  circle  and  one  such  ray.  The  origin  itself  is  special:  No  circle 
passes  through  it,  and  all  the  rays  begin  at  it.  (See  Figure  1 .84.)  For  points  P  other 
than  the  origin,  we  assign  to  P  the  polar  coordinates  (r,  9),  where  r  is  the  radius 
of  the  circle  on  which  P  lies  and  9  is  the  angle  between  the  positive  x-axis  and 
the  ray  on  which  P  lies.  (9  is  measured  as  opening  counterclockwise.)  The  origin 
is  an  exception:  It  is  assigned  the  polar  coordinates  (0,  9),  where  9  can  be  any 
angle.  (See  Figure  1.85.)  As  we  have  described  polar  coordinates,  r  >  0  since  r 
is  the  radius  of  a  circle.  It  also  makes  good  sense  to  require  0  <  9  <  2jt,  for  then 
every  point  in  the  plane,  except  the  origin,  has  a  uniquely  determined  pair  of  polar 
coordinates.  Occasionally,  however,  it  is  useful  not  to  restrict  r  to  be  nonnegative 
and  9  to  be  between  0  and  lit.  In  such  a  case,  no  point  of  R2  will  be  described  by 
a  unique  pair  of  polar  coordinates:  If  P  has  polar  coordinates  (r,  9),  then  it  also 
has  (r,  9  +  2nn)  and  (— r,  9  +  (2n  +  \)tt)  as  coordinates,  where  n  can  be  any 
integer.  (To  locate  the  point  having  coordinates  (r,  9),  where  r  <  0,  construct  the 
ray  making  angle  9  with  respect  to  the  positive  x-axis,  and  instead  of  marching 
|  r  |  units  away  from  the  origin  along  this  ray,  go  |  r  \  units  in  the  opposite  direction, 
as  shown  in  Figure  1.86.) 

EXAMPLE  1  Polar  coordinates  may  already  be  familiar  to  you.  Nonetheless, 
make  sure  you  understand  that  the  points  pictured  in  Figure  1.87  have  the  coor- 
dinates indicated.  ♦ 


(3V3,  kI6) 


(2,  5n/6) 

(2,  kI6) 

(5,0) 

(-1,  5tt/6)  or 

(1, 1171/6)  or 

(3,  3tt/2)  < 

(l,-7i/6) 

Figure  1 .87  Figure  for 
Example  1. 


Figure  1 .88  The  graph  of 
r  =  6  cos  6  in  Example  2. 


EXAMPLE  2  Let's  graph  the  curve  given  by  the  polar  equation  r  =  6cos# 
(Figure  1.88).  We  can  begin  to  get  a  feeling  for  the  graph  by  compiling  values,  as 
in  the  adjacent  tabulation. 


64       Chapter  1  [  Vectors 

Thus,  r  decreases  from  6  to  0  as  9  increases  from  0  to  ji/2;  r  decreases  from 
0  to  —6  (or  is  not  defined  if  you  take  r  to  be  nonnegative)  as  9  varies  from  ir/2 
to  7r;  r  increases  from  —6  to  0  as  9  varies  from  n  to  3jt/2;  and  r  increases  from 
0  to  6  as  9  varies  from  3jt/2  to  2ir .  To  graph  the  resulting  curve,  imagine  a  radar 
screen:  As  9  moves  counterclockwise  from  0  to  2n,  the  point  (r,  9)  of  the  graph  is 
traced  as  the  appropriate  "blip"  on  the  radar  screen.  Note  that  the  curve  is  actually 
traced  twice:  once  as  9  varies  from  0  to  n  and  then  again  as  9  varies  from  it  to 
2jt.  Alternatively,  the  curve  is  traced  just  once  if  we  allow  only  9  values  that  yield 
nonnegative  r  values.  The  resulting  graph  appears  to  be  a  circle  of  radius  3  (not 
centered  at  the  origin),  and  in  fact,  one  can  see  (as  in  Example  3)  that  the  graph 
is  indeed  such  a  circle.  ♦ 


The  basic  conversions  between  polar  and  Cartesian  coordinates  are  provided 

by  the  following  relations: 

Polar  to  Cartesian: 

J  x  =  r  cos9  , 
1  y  =  r  sin  9  ' 

Cartesian  to  polar: 

\  r2  =  x2  +  y2 

(tan<9  =  y/x   -  (2) 

Note  that  the  equations  in  (2)  do  not  uniquely  determine  r  and  9  in  terms  of  x 
and  y.  This  is  quite  acceptable,  really,  since  we  do  not  always  want  to  insist  that 
r  be  nonnegative  and  9  be  between  0  and  2n.  If  we  do  restrict  r  and  9,  however, 
then  they  are  given  in  terms  of  x  and  y  by  the  following  formulas: 

r  =  y/x2  +  y2, 


tan-1  y/x 

if  x 

> 

0,y 

> 

0 

tan-1  y/x  +  2jt 

ifx 

> 

0,y 

< 

0 

tan-1  y/x  +  n 

ifx 

< 

0,y 

> 

0 

tt/2 

ifx 

0,y 

> 

0 

3tt/2 

ifx 

0,y 

< 

0 

indeterminate 

ifx 

y  = 

0 

The  complicated  formula  for  9  arises  because  we  require  0  <  9  <  2n ,  while  the 
inverse  tangent  function  returns  values  between  —  ir/2  and  n/2  only.  Now  you 
see  why  the  equations  given  in  (2)  are  a  better  bet! 

EXAMPLE  3  We  can  use  the  formulas  in  (1)  and  (2)  to  prove  that  the  curve  in 
Example  2  really  is  a  circle.  The  polar  equation  r  =  6  cos  9  that  defines  the  curve 
requires  a  little  ingenuity  to  convert  to  the  corresponding  Cartesian  equation.  The 
trick  is  to  multiply  both  sides  of  the  equation  by  r.  Doing  so,  we  obtain 

r2  =  6r  cosO. 

Now  (1)  and  (2)  immediately  give 

x2  +  y2  =  6x. 


1.7  |  New  Coordinate  Systems  65 
We  complete  the  square  in  x  to  find  that  this  equation  can  be  rewritten  as 

(x  -  3)2  +  y2  =  9, 

which  is  indeed  a  circle  of  radius  3  with  center  at  (3,  0).  ♦ 


Cylindrical  Coordinates  

Cylindrical  coordinates  on  R3  are  a  "naive"  way  of  generalizing  polar  coordinates 
to  three  dimensions,  in  the  sense  that  they  are  nothing  more  than  polar  coordinates 
used  in  place  of  the  x-  and  y-coordinates.  (The  z-coordinate  is  left  unchanged.) 
The  geometry  is  as  follows:  Except  for  the  z-axis,  fill  all  of  space  with  infinitely 
extended  circular  cylinders  with  axes  along  the  z-axis  as  in  Figure  1.89.  Then 
any  point  P  in  R3  not  lying  on  the  z-axis  lies  on  exactly  one  such  cylinder. 
Hence,  to  locate  such  a  point,  it's  enough  to  give  the  radius  of  the  cylinder,  the 
circumferential  angle  9  around  the  cylinder,  and  the  vertical  position  z  along  the 
cylinder.  The  cylindrical  coordinates  of  P  are  (r,  9,  z),  as  shown  in  Figure  1.90. 
Algebraically,  the  equations  in  (1)  and  (2)  can  be  extended  to  produce  the  basic 
conversions  between  Cartesian  and  cylindrical  coordinates. 


The  basic  conversions  between  cylindrical 

and  Cartesian  coordinates 

are 

provided  by  the  following  relations: 

x  =  r  cos  6* 

Cylindrical  to  Cartesian: 

y  =  r  sin  9  ; 

Z  =  Z 

(3) 

r2  =  x2  +  y2 

Cartesian  to  cylindrical: 

tan  9  =  y/x  . 

(4) 

z  =  z 

As  with  polar  coordinates,  if  we  make  the  restrictions  r  >  0,  0  <  9  <  2n,  then 
all  points  of  R3  except  the  z-axis  have  a  unique  set  of  cylindrical  coordinates.  A 
point  on  the  z-axis  with  Cartesian  coordinates  (0,0,  zo)  has  cylindrical  coordinates 
(0,  9,  zo),  where  9  can  be  any  angle. 

Cylindrical  coordinates  are  useful  for  studying  objects  possessing  an  axis  of 
symmetry.  Before  exploring  a  few  examples,  let's  understand  the  three  "constant 
coordinate"  surfaces. 

•  The  r  =  T(,  surface  is,  of  course,  just  a  cylinder  of  radius  ro  with  axis  the 
z-axis.  (See  Figure  1.91.) 

•  The  9  =  9q  surface  is  a  vertical  plane  containing  the  z-axis  (or  a  half-plane 
with  edge  the  z-axis  if  we  take  r  >  0  only).  (See  Figure  1.92.) 

•  The  z  =  zo  surface  is  a  horizontal  plane.  (See  Figure  1.93.) 


Half-plane  only 


EXAMPLE  4  Graph  the  surface  having  cylindrical  equation  r  =  6  cos  8.  (This 
equation  is  identical  to  the  one  in  Example  2.)  In  particular,  z  does  not  appear 
in  this  equation.  What  this  means  is  that  if  the  surface  is  sliced  by  the  horizontal 
plane  z  =  c  where  c  is  a  constant,  we  will  see  the  circle  shown  in  Example  2,  no 
matter  what  c  is.  If  we  stack  these  circular  sections,  then  the  entire  surface  is  a 
circular  cylinder  of  radius  3  with  axis  parallel  to  the  z-axis  (and  through  the  point 
(3,  0,  0)  in  cylindrical  coordinates).  This  surface  is  shown  in  Figure  1.94.  ♦ 

EXAMPLE  5  Graph  the  surface  having  equation  z  =  2r  in  cylindrical  co- 
ordinates. 

Here  the  variable  6  does  not  appear  in  the  equation,  which  means  that  the 
surface  in  question  will  be  circularly  symmetric  about  the  z-axis.  In  other  words, 
if  we  slice  the  surface  by  any  plane  of  the  form  6  =  constant  (or  half-plane,  if  we 
take  r  >  0),  we  see  the  same  curve,  namely,  a  line  (respectively,  a  half-line)  of 
slope  2.  As  we  let  the  constant-0  plane  vary,  this  line  generates  a  cone,  as  shown 
in  Figure  1 .95.  The  cone  consists  only  of  the  top  half  (nappe)  when  we  restrict  r 
to  be  nonnegative. 

The  Cartesian  equation  of  this  cone  is  readily  determined.  Using  the  formulas 
in  (4),  we  have 

Z  =  2r     =>     z2=4r2     <^=>     z2  =  4(x2  +  y2). 

Since  z  can  be  positive  as  well  as  negative,  this  last  Cartesian  equation  describes 
the  cone  with  both  nappes.  If  we  want  the  top  nappe  only,  then  the  equation 
z  =  2y 'x2  +  y2  describes  it.  Similarly,  z  =  —2^x2  +  y2  describes  the  bottom 
nappe.  ♦ 

Spherical  Coordinates   

Fill  all  of  space  with  spheres  centered  at  the  origin  as  in  Figure  1 .96.  Then  every 
point  P  e  R3,  except  the  origin,  lies  on  a  single  such  sphere.  Roughly  speaking, 
the  spherical  coordinates  of  P  are  given  by  specifying  the  radius  p  of  the  sphere 
containing  P  and  the  "latitude  and  longitude"  readings  of  P  along  this  sphere. 
More  precisely,  the  spherical  coordinates  (p,  <p,  9)  of  P  are  defined  as  follows:  p 
is  the  distance  from  P  to  the  origin;  <p  is  the  angle  between  the  positive  z-axis  and 
the  ray  through  the  origin  and  P;  and  6  is  the  angle  between  the  positive  x-axis 
and  the  ray  made  by  dropping  a  perpendicular  from  P  to  the  xy-plane.  (See  Figure 
1 .97.)  The  0 -coordinate  is  exactly  the  same  as  the  ^-coordinate  used  in  cylindrical 
coordinates.  (Warning:  Physicists  usually  prefer  to  reverse  the  roles  of  <p  and  9, 
as  do  some  graphical  software  packages.) 


1.7  |  New  Coordinate  Systems  67 


z 


Figure  1 .96  The  spherical  Figure  1 .97  Locating  the  point 

coordinate  system.  P,  using  spherical  coordinates. 


It  is  standard  practice  to  impose  the  following  restrictions  on  the  range  of 
values  for  the  individual  coordinates: 

P  >  0,       0  <  <p  <  n,       0  <  9  <  lit.  (5) 

With  such  restrictions,  all  points  of  R3 ,  except  those  on  the  z-axis,  have  a  uniquely 
determined  set  of  spherical  coordinates.  Points  along  the  z-axis,  except  for  the 
origin,  have  coordinates  of  the  form  (po,  0,  9)  or  (po,  n,  9),  where  po  is  a  positive 
constant  and  9  is  arbitrary.  The  origin  has  spherical  coordinates  (0,<p,9),  where 
both  (p  and  6  are  arbitrary. 


EXAMPLE  6  Several  points  and  their  corresponding  spherical  coordinates  are 
shown  in  Figure  1.98.  ♦ 


Figure  1.98  Figure  for  Example  6.  Figure  1.99  The  graph  of  Figure  1.100  The  spherical 

P  =  Po  (>  0).  surface  <p  =  (po,  shown  for 

different  values  of  (po. 

Spherical  coordinates  are  especially  useful  for  describing  objects  that  have 
a  center  of  symmetry.  With  the  restrictions  given  by  the  inequalities  in  (5),  the 
constant  coordinate  surface  p  =  po  (po  >  0)  is,  of  course,  a  sphere  of  radius  po, 
as  shown  in  Figure  1.99.  The  surface  given  by  9  =  9o  is  a  half-plane  just  as  in  the 
cylindrical  case.  The  <p  =  <po  surface  is  a  single-nappe  cone  if  cpo  /  tt/2  and  is 
the  xy-plane  if  cp0  =  tt/2  (and  is  the  positive  or  negative  z-axis  if  (po  =  0  or  it). 
(See  Figure  1 . 1 00.)  If  we  do  not  insist  that  p  be  nonnegative,  then  the  cones  would 
include  both  nappes. 


68       Chapter  1  [  Vectors 


The  basic  equations  relating  spherical  coordinates  to  both  cylindrical  and 

Cartesian  coordinates  are  as  follows. 

Spherical/cylindrical: 

r  =  p  sin  cp 

—2         2   i  _2 

p  =  r  +  z 

9  =  9 

\my  =  r/z    .  (6) 

z  =  p  cos  cp 

9=9 

Spherical/Cartesian: 

x  =  psincp  cos 9 

p2  =  x2  +  y2  +  z2 

y  =  p  sin  (p  sin  9 

tan<p  =  y/x2  +  y2/z-  (7) 

z  =  p  coscp 

tan#  =  y/x 

Figure  1.101  Converting 
spherical  to  cylindrical  coordinates 
when  0  <  <p  <  j. 


Figure  1.102  Converting 
spherical  to  cylindrical 
coordinates  when  jt/2  <  cp  <  it . 


Using  basic  trigonometry,  it  is  not  difficult  to  establish  the  conversions  in  (6). 
From  the  right  triangle  shown  in  Figure  1.101,  we  have 


COS  I 


Hence, 
Similarly, 

so  that 


r  =  p  cos 


p  sin<p. 


sin 


p  sin 


(*-')- 

(£-')- 


z 
P 

p  cos  (p. 


Thus,  the  formulas  in  (6)  follow  when  0  <  <p  <  tt/2.  If  it/2  <  <p  <  n,  then  we 
may  employ  Figure  1.102.  So 


p  cos 


('-I) 


p  sm<p, 


and 


z  =  -P  sin  ^  -  -)  =  p  sin       -  <p)  = 


p  cos<p. 


1.7  |  New  Coordinate  Systems  69 


V 

p  =  2a  cos  <p 

0 

2a 

7T/6 

V3a 

ji/4 

V2a 

tt/3 

a 

JT/2 

0 

2jt/3 

—a 

3jt/4 

— \/2a 

7T 

—2a 

Hence,  the  relations  in  (6)  hold  in  general.  The  equations  in  (7)  follow  by  substi- 
tution of  those  in  (6)  into  those  of  (3)  and  (4). 

EXAMPLE  7  The  cylindrical  equation  z  =  2r  in  Example  5  converts  via  (6) 
to  the  spherical  equation 

p  cosco  =  2p  simp. 

Therefore, 


tano5 


26°. 


1  -i  1 
-     <=?■     co  =  tan  - 

2  r  2 

Thus,  the  equation  defines  a  cone  (as  we  just  saw).  The  spherical  equation  is 
especially  simple  in  that  it  involves  just  a  single  coordinate.  ♦ 


EXAMPLE  8   Not  all  spherical  equations  are  improvements  over  their  cylindri 

y 


cal  or  Cartesian  counterparts.  For  example,  the  Cartesian  equation  6x  =  x2  1 


(whose  polar-cylindrical  equivalent  is  r  =  6  cos  9)  becomes 

6p  sin  (p  cos  9  =  p2  sin2  05  cos2  9  +  p2  sin2  cp  sin2  9 
from  (7).  Simplifying, 


6p  sin  (p  cos  9  =  p  sin  <p  (cos  ©  +  sin  9) 
6p  sin  09  cos  9  =  p2  sin2  <p 
6cos#  =  p  sin  05. 


This  spherical  equation  is  more  complicated  than  the  original  Cartesian  equation 
in  that  all  three  spherical  coordinates  are  involved.  Therefore,  it  is  not  at  all 
obvious  that  the  spherical  equation  describes  a  cylinder.  ♦ 

EXAMPLE  9  Let's  graph  the  surface  with  spherical  equation  p  =  2a  cosoo, 
where  a  >  0.  As  with  the  graph  of  the  cone  with  cylindrical  equation  z  =  2r, 
note  that  the  equation  is  independent  of  9.  Thus,  all  sections  of  this  surface  made 
by  slicing  with  the  half-plane  9  =  c  must  be  the  same.  If  we  compile  values  as 
in  the  adjacent  table,  then  the  section  of  the  surface  in  the  half-plane  9  =  0  is  as 
shown  in  Figure  1.103.  Since  this  section  must  be  identical  in  all  other  constant-0 
half-planes,  we  see  that  this  surface  appears  to  be  a  sphere  of  radius  a  tangent  to 
the  xy-plane,  which  is  shown  in  Figure  1.104. 


(2a,  0, 0) 

Section  of 
p  =  2a  cos  q> 


Figure  1.103  The  cross  section 
of  p  =  2a  cos  05  in  the  half-plane 
6  =  0. 


Figure  1 .1 04  The  graph  of  p  = 
2a  cos<p. 


Chapter  1  |  Vectors 


The  Cartesian  equation  of  the  surface  is  determined  by  multiplying  both  sides 
of  the  spherical  equation  by  p  and  using  the  conversion  equations  in  (7): 

p  =  2a  cos  <p         p2  =  lap  cos  <p 

<^=>  x2  +  y2  +  z2  =  2az 
<^=>  x2  +  y2  +  (z  -  a)2  =  a2 

by  completing  the  square  in  z.  This  last  equation  can  be  recognized  as  that  of  a 
sphere  of  radius  a  with  center  at  (0,  0,  a)  in  Cartesian  coordinates.  ♦ 

EXAMPLE  10  NASA  launches  a  10-ft-diameter  space  probe.  Unfortunately, 
a  meteor  storm  pushes  the  probe  off  course,  and  it  is  partially  embedded  in 
the  surface  of  Venus,  to  a  depth  of  one  quarter  of  its  diameter.  To  attempt  to 
reprogram  the  probe's  on-board  computer  to  remove  it  from  Venus,  it  is  necessary 
to  describe  the  embedded  portion  of  the  probe  in  spherical  coordinates.  Let  us 
find  the  description  desired,  assuming  that  the  surface  of  Venus  is  essentially  flat 
in  relation  to  the  probe  and  that  the  origin  of  our  coordinate  system  is  at  the  center 
of  the  probe. 


z  =  -5/2 

Figure  1 .1 07  Coordinate  view  of 
the  cross  section  of  the  probe  of 
Example  10. 


The  situation  is  illustrated  in  Figure  1 . 1 05 .  The  buried  part  of  the  probe  clearly 
has  symmetry  about  the  z-axis.  That  is,  any  slice  by  the  half-plane  9  =  constant 
looks  the  same  as  any  other.  Thus,  9  can  vary  between  0  and  2it .  A  typical  slice 
of  the  probe  is  shown  in  Figure  1 . 106.  Elementary  trigonometry  indicates  that  for 
the  angle  a  in  Figure  1.106, 

1  1 
cos  a  =  —  =  -. 
5  2 

Hence,  a  =  cos-1  \  =  n/3.  Thus,  the  spherical  angle  <p  (which  opens  from  the 
positive  z-axis)  varies  from  n  —  tt/3  =  2n/3  to  it  as  it  generates  the  buried  part 
of  the  probe.  Finally,  note  that  for  a  given  value  of  y>  between  2n/3  and  it,  p 
is  bounded  by  the  surface  of  Venus  (the  plane  z  =  —  |  in  Cartesian  coordinates) 
and  the  spherical  surface  of  the  probe  (whose  equation  in  spherical  coordinates  is 
p  =  5).  See  Figure  1.107.  From  the  formulas  in  (7)  the  equation  z  =  —  f  corre- 
sponds to  the  spherical  equation  p  cos  (p  =  —  |  or,  equivalently,  to  p  =  —  |  sec  <p. 
Therefore,  the  embedded  part  of  the  probe  may  be  defined  by  the  set 


5  2tt  „  ' 

-  seccp  <  p  <  5,  —  <(p<tt,0<9<2tt 


1.7  |  New  Coordinate  Systems 


Standard  Bases  for  Cylindrical 

and  Spherical  Coordinates   

In  Cartesian  coordinates,  there  are  three  special  unit  vectors  i,  j,  and  k  that  point 
in  the  directions  of  increasing  x-,  y-,  and  z-coordinate,  respectively.  We  find 
corresponding  sets  of  vectors  for  cylindrical  and  spherical  coordinates.  That  is, 
in  each  set  of  coordinates,  we  seek  mutually  orthogonal  unit  vectors  that  point  in 
the  directions  of  increasing  coordinate  values. 

In  cylindrical  coordinates,  the  situation  is  as  shown  in  Figure  1.108.  The 
vectors  e, ,  ee,  and  e-,  which  form  the  standard  basis  for  cylindrical  coordinates, 
are  unit  vectors  that  each  point  in  the  direction  in  which  only  the  coordinate 
indicated  by  the  subscript  increases.  There  is  an  important  difference  between 
the  standard  basis  vectors  in  Cartesian  and  cylindrical  coordinates.  In  the  former 
case,  i,  j,  and  k  do  not  vary  from  point  to  point.  However,  the  vectors  e,  and  eg 
do  change  as  we  move  from  point  to  point. 

Now  we  give  expressions  for  er,  eg,  and  ez.  Since  the  cylindrical  z-coordinate 
is  the  same  as  the  Cartesian  z-coordinate,  we  must  have  ez  =  k.  The  vector 
er  must  point  radially  outward  from  the  z-axis  with  no  k-component.  At  a 
point  (x,  y,  z)  6  R3  (Cartesian  coordinates),  the  vector  x'\  +  y\  has  this  prop- 
erty. Normalizing  it  to  obtain  a  unit  vector  (see  Proposition  3.4  of  §1.3),  we 
obtain 

=   xi  +  y\ 

v7*2  +  y2 ' 

With  er  and  e-  in  hand  it's  now  a  simple  matter  to  define  eg,  since  it  must  be 
perpendicular  to  both  e,-  and  ez.  We  take 

-yi  +  xj 

eg  =  e;xer  = 

V*2  +  y2 

(The  reason  for  this  choice  of  cross  product,  as  opposed  to  e,  x  ez,  is  so  that  eg 
points  in  the  direction  of  increasing  0.)  To  summarize,  and  using  the  cylindrical 
to  Cartesian  conversions  given  in  (3), 


e, 

xi  +  yj 

=  —  =  cos#  i  +  sm6  j; 

V*2  +  y2 

eg 

-yi  +  *j       .  a.  .  a. 

=  —  =  —  sin0 1  +  cos0  j; 

V*2  +  v2 

(8) 

=  k. 

In  spherical  coordinates,  the  situation  is  shown  in  Figure  1 . 109.  In  particular, 
there  are  three  unit  vectors  ep ,  ,  and  eg  that  form  the  standard  basis  for  spherical 
coordinates.  These  vectors  all  change  direction  as  we  move  from  point  to  point. 

We  give  expressions  for  ep ,  ,  and  ee .  Since  the  ^-coordinates  in  both  spher- 
ical and  cylindrical  coordinates  mean  the  same  thing,  eg  in  spherical  coordinates 
is  given  by  the  value  of  eg  in  (8).  At  a  point  (x,y,  z),  the  vector  ep  should  point 


72       Chapter  1  [  Vectors 

from  the  origin  directly  to  (x,  y,  z).  Thus,  ep  may  be  obtained  by  normalizing 
■*i  +  yi  +  zk.  Finally,  ep  is  nothing  more  than  eg  x  e^.  If  we  explicitly  perform 
the  calculations  just  described  and  make  use  of  the  conversion  formulas  in  (7), 
the  following  are  obtained: 


xi  +  yj  +  zk 

=  —  =  sm<p  cos  6 

Vx2  +  y2  +  z2 

i  +  sin  <p  sin  0  j  +  cos  <p  k; 

xzi  +  yzj  -  (x2  +  y2)k 

v7*2  +  y2  Vx2  +  y2  +  ^2 

(9) 

sin  <p  k; 

=  cos  <p  cos  0  i  +  cos  cp  sin  0  j  - 

— yi  +  xj 
=  —  =  —  sm  6 1  +  cos  6 

V*2  + 

Although  the  results  of  (8)  and  (9)  will  not  be  used  frequently,  they  will  prove 
helpful  on  occasion. 

Hyperspherical  Coordinates  (optional) 

There  is  a  way  to  provide  a  set  of  coordinates  for  R"  that  generalizes  spherical 
coordinates  on  R3.  For  n  >  3,  the  hyperspherical  coordinates  of  a  point  P  e  R" 
are  (p,  q>\,  q>%, . . . ,  <pn-i)  and  are  defined  by  their  relations  with  the  Cartesian 
coordinates  (xi,  x%, . . . ,  xn)  of  P  as 

'  xi  =  p  sin     sin  ip2  ■  ■  ■  sin  <p„_2  cos  <p„_  i 

x2  =  p  sin^i  sin  ip2  ■  ■  ■  sin<p„_2  sin<pn_i 

x3  =  p  sinip!  sin     •  •  •  sin<p,,_3  cos<p„_2 

(10) 

X4  =  p  sm  ipi  sm  q>2  -  -  ■  sm  <p„_4  cos  <p„_3 


_xn  =  p  COS^i 
To  be  more  explicit,  in  equation  (10)  above  we  take 

X*  =  p  sin     sin  cp2  -  ■  ■  sin        cos  ^„_jt+i    for  k  =  3, . . . ,  n. 
Note  that  when  n  =  3,  the  relations  in  (10)  become 

xi  =  p  sin^i  cos  i^2 
•  X2  =  p  sin^i  sini£>2  ■ 
*3  =  pcos^i 

These  relations  are  the  same  as  those  given  in  (7),  so  hyperspherical  coordinates 
are  indeed  the  same  as  spherical  coordinates  when  n  =  3. 

In  analogy  with  (5),  it  is  standard  practice  to  impose  the  following  restrictions 
on  the  range  of  values  for  the  coordinates: 

P  >  0,       0  <  (pi  <  7T  for  k  =  1, . . . ,  n  —  2,       0  <  q>n-\  <  lit.     (1 1) 


1.7  |  Exercises  73 


Then,  with  these  restrictions,  we  can  convert  from  hypersphencal  coordinates  to 
Cartesian  coordinates  by  means  of  the  following  formulas: 

'  p2  =  x\  +  x\  H  hi„2 

tanpi  =  ^2  H  hx2_,/x„ 

tan  ^2  =  J  A  H  h  x*_2/Xn-i  n  9, 


tan<p„_2  =  J     +  *f/*3 
tan^„_!  =  x2/*i 

Hyperspherical  coordinates  get  their  name  from  the  fact  that  the  (n  —  1)- 
dimensional  hypersurface  in  R"  defined  by  the  equation  p  =  p0,  where  po  is  a 
positive  constant,  consists  of  points  on  the  hypersphere  of  radius  po  defined  in 
Cartesian  coordinates  by  the  equation 

x2+x2  +  ---+x2  =  p2. 


1.7  Exercises 


In  Exercises  1—3,  find  the  Cartesian  coordinates  of  the  points 
whose  polar  coordinates  are  given. 

1.  (V2,jr/4) 

2.  (V3,  5jt/6) 

3.  (3,  0) 

In  Exercises  4-6,  give  a  set  of polar  coordinates  for  the  point 
whose  Cartesian  coordinates  are  given. 

4.  (273,2) 

5.  (-2,2) 

6.  (-1,-2) 

In  Exercises  7-9,  find  the  Cartesian  coordinates  of  the  points 
whose  cylindrical  coordinates  are  given. 

7.  (2,2,2) 

8.  (n,n/2,  1) 

9.  (1,2* /3,  -2) 

In  Exercises  10-13,  find  the  rectangular  coordinates  of  the 
points  whose  spherical  coordinates  are  given. 

10.  (4,  ;r/2,7r/3) 

11.  (3,  ;r/3,7r/2) 

12.  (1,3tt/4,  2jt/3) 

13.  (2,;r,7r/4) 


In  Exercises  14-16,  find  a  set  of  cylindrical  coordinates  of  the 
point  whose  Cartesian  coordinates  are  given. 

14.  (-1,0,2) 

15.  (-1,  V3,  13) 

16.  (5,6,3) 

In  Exercises  1 7  and  18,  find  a  set  of  spherical  coordinates  of 
the  point  whose  Cartesian  coordinates  are  given. 

17.  (1,  -1,  v/6) 

18.  (0,  V3,  1) 

19.  This  problem  concerns  the  surface  described  by  the 
equation  (r  —  2)2  +  z2  =  1  in  cylindrical  coordinates. 
(Assume  r  >  0.) 

(a)  Sketch  the  intersection  of  this  surface  with  the  half- 
plane  8  =  jt/2. 

(b)  Sketch  the  entire  surface. 

20.  (a)  Graph  the  curve  in  R2  having  polar  equation  r  = 

2a  sin  9,  where  a  is  a  positive  constant. 

(b)  Graph  the  surface  in  R3  having  spherical  equation 
p  =  2a  sin  <p. 

21.  Graph  the  surface  whose  spherical  equation  is  p  = 
1  —  cosip. 

22.  Graph  the  surface  whose  spherical  equation  is  p  = 
1  —  sinip. 

In  Exercises  23-25,  translate  the  following  equations  from 
the  given  coordinate  system  (i.e.,  Cartesian,  cylindrical,  or 


74       Chapter  1  |  Vectors 


spherical)  into  equations  in  each  of  the  other  two  systems.  In 
addition,  identify  the  surfaces  so  described  by  providing  ap- 
propriate sketches. 

23.  p  sin<p  sin#  =  2 

24.  z2  =  2x2  +  2y2 

25.  r  =  0 

In  Exercises  26-29,  sketch  the  solid  whose  cylindrical  coordi- 
nates (r,  9,  z)  satisfy  the  given  inequalities. 

26.  0<r<3,     0<6»<tt/2,     -1  <  z  <  2 

27.  r  <  z  <  5,     0  <  9  <  n 

28.  2r  <  z  <  5  -  3r 

29.  r2  -  1  <  z  <  5  -  r2 

In  Exercises  30-35,  sketch  the  solid  whose  spherical  coordi- 
nates (p,  <p,  9)  satisfy  the  given  inequalities. 

30.  1  <  p  <  2 

31.  0  <  p  <  1,    0  <  (p  <  tt/2 

32.  0  <  p  <  1,    0  <  6»  <  jr/2 

33.  0  <  p  <  jt/4,    0  <  p  <  2 

34.  0  <  p  <  2/coscp,    0  <  tp  <  tt/4 

35.  2  cos  (p  <  p  <3 

36.  (a)  Which  points  P  in  R2  have  the  same  rectangular 

and  polar  coordinates? 

(b)  Which  points  P  in  R3  have  the  same  rectangular 
and  cylindrical  coordinates? 

(c)  Which  points  P  in  R3  have  the  same  rectangular 
and  spherical  coordinates? 

37.  (a)  How  are  the  graphs  of  the  polar  equations  r  =  f(9) 

andr  =  -f{9)  related? 

(b)  How  are  the  graphs  of  the  spherical  equations 
p  =  f((p,  9)  and  p  =  -f(q>,  0)  related? 

(c)  Repeat  part  (a)  for  the  graphs  of  r  =  f{9)  and 
r  =  3/(0). 

(d)  Repeat  part  (b)  for  the  graphs  of  p  =  f((p,  9)  and 
p  =  3f(<p,0). 

38.  Suppose  that  a  surface  has  an  equation  in  cylindrical 
coordinates  of  the  form  z  =  f(r).  Explain  why  it  must 
be  a  surface  of  revolution. 

39.  (a)  Verify  that  the  basis  vectors  er,  eg,  and  ez  for 

cylindrical  coordinates  are  mutually  perpendicu- 
lar unit  vectors. 

(b)  Verify  that  the  basis  vectors  ep ,  ,  and  eg  for  spher- 
ical coordinates  are  mutually  perpendicular  unit 
vectors. 


40.  Use  the  formulas  in  (8)  to  express  i,  j,  k  in  terms  of  er, 
eg ,  and  ez . 

41 .  Use  the  formulas  in  (9)  to  express  i,  j,  k  in  terms  of  ep, 
ev,  and  eg. 

42.  Consider  the  solid  in  R3  shown  in  Figure  1.110. 

(a)  Describe  the  solid,  using  spherical  coordinates. 

(b)  Describe  the  solid,  using  cylindrical  coordinates. 

A  portion  of  the 
z    sphere  of  radius  3 
(centered  at  origin) 


y 


x 


Figure  1.110  The  ice-cream- 
cone-like  solid  in  R3  in  Exercise  42. 

In  Exercises  43-47,  you  will  use  the  equations  in  (10)  to  es- 
tablish those  in  (12). 

43.  Show  that  tan <pn-\  =  x%lx\. 

44.  (a)  Calculate  x\  +  x\  in  terms  of  the  hyperspherical 

coordinates  p,  (pi, . . . ,  (pn-%. 

(b)  Assuming  the  inequalities  in  (1 1),  use  part  (a)  to 
show  that  tamp„_2  =  ^jx\  +  x2/x^. 

45.  (a)  Calculate  x\  +  x\  +  x2  in  terms  of  the  hyperspher- 

ical coordinates  p,  <p\, . .. ,  <p„-z. 

(b)  Assuming  the  inequalities  in  (1 1),  use  part  (a)  to 
show  that  tan  ip,,- 3  =  J 1  x\  +  x\  +  x2/x4. 

46.  (a)  For  k  —  2, . . . ,  n  —  1,  show  that  x\  +  x\  +  ■  ■  ■  + 

x\  =  p2  sin2  <p\  •  •  ■  sin2  (p„-k-  (Note:  This  is  best 
accomplished  by  means  of  mathematical  induc- 
tion.) 

(b)  Assuming  the  inequalities  in  (11),  use  part  (a) 
to  show  that,  for  k  =  2  n  —  1,  tan^„_j.  = 

jx2  +  ---+xt/xk+l. 

47.  Show  that  x\  +  x\  H  \-x2  =  p2. 


Miscellaneous  Exercises  for  Chapter  1 


True/False  Exercises  for  Chapter  1 


1.  If  a  =  (1,  7,  -9)  and  b  =  (1,  -9,  7),  then  a  =  b. 

2.  If  a  and  b  are  two  vectors  in  R3  and  k  and  /  are  real 
numbers,  then  (k  —  Z)(a  +  b)  =  ka  —  /a  +  kb  —  lb. 

3.  The  displacement  vector  from  Pi  (1,0,  —  1)  to 
P2(5,3,2)  is  (-4,-3,-3). 

4.  Force  and  acceleration  are  vector  quantities. 

5.  Velocity  and  speed  are  vector  quantities. 

6.  Displacement  and  distance  are  scalar  quantities. 

7.  If  a  particle  is  at  the  point  (2,  —1)  in  the  plane  and 
moves  from  that  point  with  velocity  vector  v  =  (1,  3), 
then  after  2  units  of  time  have  passed,  the  particle  will 
be  at  the  point  (5,  1). 

8.  The  vector  (2,3,  —2)  is  the  same  as  2i  +  3j  —  2k. 

9.  A  set  of  parametric  equations  for  the  line  through 
(1,  -2,  0)  that  is  parallel  to  (-2,  4,  7)  is  x  =  1  -  2t, 

y  =  4t-2,z  =  l. 

10.  A  set  of  parametric  equations  for  the  line  through 
(1,  2,  3)  and  (4,  3, 2)  is  x  =  4  -  3f,  y  =  3  -  t,  z  = 
t  +  2. 

1 1 .  The  line  with  parametric  equations  x  =  2  — 
3t,  y  =  t+l,  z  =  2t  —  3  has  symmetric  form 
x+2  _  _l_z-3 

-3    ~y~    ~  2 

12.  The  two  sets  of  parametric  equations  x  =  3t  —  1, 
y  =  2  -  t,  z  =  2t  +  5  and  x  =  2  -  6t,  y  =  2t  +  1, 
z  =  1  —  At  both  represent  the  same  line. 

13.  The  parametric  equations  x  =  2sinf,  y  =  2cos?, 
where  0  <  t  <  n,  describe  a  circle  of  radius  2. 

1 4.  The  dot  product  of  two  unit  vectors  is  1 . 

15.  For  any  vector  a  in  R"  and  scalar  k,  we  have  ||£a||  = 
fc||a||. 

16.  If  a,  u  e  R"  and  ||u||  =  1,  then  projua  =  (a  •  u)u. 

17.  For  any  vectors  a,  b,  c  in  R3,  we  have  a  x  (b  x  c)  = 
(a  x  b)  x  c. 


1 8.  The  volume  of  a  parallelepiped  determined  by  the  vec- 
tors a,  b,  c  e  R3  is  |(a  x  c)  •  b|. 

19.  ||a||b  —  ||b||a  is  a  vector. 

20.  (a  x  b)  •  c  —  (a  x  c)  •  b  is  a  scalar. 

21.  The  plane  containing  the  points  (1,  2,  1),  (3,  —1,  0), 
and  (1,  0,  2)  has  equation  5a:  +  2y  +  4z  =  13. 

22.  The  plane  containing  the  points  (1,  2,  1),  (3,  —1,  0), 
and  (1,0,2)  is  given  by  the  parametric  equations 

x  =  2s,  y  =  —3s  —  2t,  z  =  t  —  s. 

23.  If  A  is  a  5  x  7  matrix  and  B  is  a  7  x  7  matrix,  then  BA 
is  a  7  x  5  matrix. 


24.  If  A 


1   2  0 

-10  2 

5   9  2 

0   8  0 


then 


-1 

2 

1 

1 

0 

3 

detA  =  2 

5 

2 

0 

+  9 

-1 

2 

1 

0 

0 

-6 

0 

0 

-6 

1 

0 

3 

-8 

1 

2 

1 

5 

2 

0 

25.  If  A  is  an  n  x  n  matrix,  then  det  (2A)  =  2  det  A. 

26.  The  surface  having  equation  r  =  4  sin  9  in  cylindrical 
coordinates  is  a  cylinder  of  radius  2. 

27.  The  surface  having  equation  p  =  4  cos  6  in  spherical 
coordinates  is  a  sphere  of  radius  2. 

28.  The  surface  having  equation  p  cos  6  sin  cp  =  3  in  spher- 
ical coordinates  is  a  plane. 

29.  The  surface  having  equation  p  =  3  in  spherical  coor- 
dinates is  the  same  as  the  surface  whose  equation  in 
cylindrical  coordinates  is  r2  +  z2  =  9. 

30.  The  surface  whose  equation  in  cylindrical  coordinates 
is  z  =  2r  is  the  same  as  the  surface  whose  equation  in 
spherical  coordinates  is  <p  =  jt/6. 


Miscellaneous  Exercises  for  Chapter  1 


1 .  If  P\ ,  Pi, . . . ,  Pn  are  the  vertices  of  a  regular  polygon 
having  n  sides  and  if  O  is  the  center  of  the  polygon, 

show  that  Y^=\  OP,  =  0.  The  case  n  =  5  is  shown  in 


Figure  1.111.  (Hint:  Don't  try  using  coordinates.  Use 
instead  sketches,  geometry,  and  perhaps  translations  or 
rotations.) 


Chapter  1  |  Vectors 


p4  P, 
Figure  1.111  The  case  n  =  5. 

2.  Find  parametric  equations  for  the  line  through  the  point 
(1,  0,  —2)  that  is  parallel  to  the  line  x  =  3t  +  1,  y  = 
5  -  7r,  z  =  ;  +  12. 

3.  Find  parametric  equations  for  the  line  through  the 
point  (1,  0,  —2)  that  intersects  the  line  x  =  3t  +  1, 
y  =  5  —  It ,  z  =  t  +  12  orthogonally.  (Hint:  Let  Xq  = 
3%  +  1,  yo  =  5  —  ltd,  zo  =  to  +  12  be  the  point  where 
the  desired  line  intersects  the  given  line.) 

4.  Given  two  points  Po(ai,  cti,  ai)  and  P\(b\,  £>2,  fo),  we 
have  seen  in  equations  (3)  and  (4)  of  §1.2  how  to 
parametrize  the  line  through  Pq  and  Pi  as  r(r)  = 
O  Pq  +  t  Pq  Pi  ,  where  t  can  be  any  real  number.  (Recall 
that  r  =  OP,  the  position  vector  of  an  arbitrary  point 
P  on  the  line.) 

(a)  For  what  value  of  /  does  r(f)  =  OPb?  For  what 
value  of  t  does  r(r)  =  OP\l 

(b)  Explain  how  to  parametrize  the  line  segment  join- 
ing Pq  and  Pi.  (See  Figure  1.112.) 


z 


y 


Figure  1.112  The  segment  joining  Po 
and  Pi  is  a  portion  of  the  line  containing 
Po  and  P|.  (See  Exercise  4.) 

(c)  Give  a  set  of  parametric  equations  for  the  line  seg- 
ment joining  the  points  (0,  1,  3)  and  (2,  5,  —7). 

5.  Recall  that  the  perpendicular  bisector  of  a  line  seg- 
ment in  R2  is  the  line  through  the  midpoint  of  the  seg- 
ment that  is  orthogonal  to  the  segment. 


(a)  Give  a  set  of  parametric  equations  for  the  perpen- 
dicular bisector  of  the  segment  joining  the  points 
Pi(-l,3)andP2(5,  -7). 

(b)  Given  general  points  Pi(ai,fl2)  and  P2(b\,b2), 
provide  a  set  of  parametric  equations  for  the  per- 
pendicular bisector  of  the  segment  joining  them. 

6.  If  we  want  to  consider  a  perpendicular  bisector  of  a 
line  segment  in  R3 ,  we  will  find  that  the  bisector  must 
be  a  plane. 

(a)  Give  an  (implicit)  equation  for  the  plane  that  serves 
as  the  perpendicular  bisector  of  the  segment  join- 
ing the  points  P(6,  3,  -2)  and  P2(-4,  1,  0). 

(b)  Given  general  points  P\{a\,  02, 03)  and 
Pi{b\,  bj,  ^3),  provide  an  equation  for  the  plane 
that  serves  as  the  perpendicular  bisector  of  the 
segment  joining  them. 

7.  Generalizing  Exercises  5  and  6,  we  may  define  the  per- 
pendicular bisector  of  a  line  segment  in  R"  to  be  the 
hyperplane  through  the  midpoint  of  the  segment  that 
is  orthogonal  to  the  segment. 

(a)  Give  an  equation  for  the  hyperplane  in  R5  that 
serves  as  the  perpendicular  bisector  of  the  seg- 
ment joining  the  points  Pi(l,  6,  0,  3,  —  2)  and 
P2(-3,-2,4,  1,0). 

(b)  Given  arbitrary  points  P\(a\ ,  . . . ,  a„)  and 
Pzibi, , . . ,  bn)  in  R",  provide  an  equation  for 
the  hyperplane  that  serves  as  the  perpendicular 
bisector  of  the  segment  joining  them. 

8.  If  a  and  b  are  unit  vectors  in  R3,  show  that 

|| a  x  b||2  +  (a-b)2  =  1. 

9.  (a)  If  a  •  b  =  a  •  c,  does  it  follow  that  b  =  c?  Explain 

your  answer. 

(b)  If  a  x  b  =  a  x  c,  does  it  follow  that  b  =  c? 
Explain. 

10.  Show  that  the  two  lines 

h  :  x  =  t-3,  y=\-2t,  z  =  2t  +  5 
h  :    x  =  4  -  It,    y  =  At  +  3,    z  =  6  -  At 

are  parallel,  and  find  an  equation  for  the  plane  that 
contains  them. 

1 1 .  Consider  the  two  planes  x  +  y  =  1  and  y  +  z  =  1 . 
These  planes  intersect  in  a  straight  line. 

( a)  Find  the  ( acute)  angle  of  intersection  between  these 
planes. 

(b)  Give  a  set  of  parametric  equations  for  the  line  of 
intersection. 

1 2.  Which  of  the  following  lines  whose  parametric  equa- 
tions are  given  below  are  parallel?  Are  any  the  same? 
(a)  x  =  At  +  6,  y  =  2  -  It,  z  =  8*  +  1 


Miscellaneous  Exercises  for  Chapter  1 


(b)  x  =  3  -  6t ,  y  =  3f ,  z  =  4  -  9t 

(c)  x  =  2  -  2t,  y  =  t  +  4,  z  =  -4r  -  7 

(d)  jc  =  2f  +  4,  y  =  1  -  f ,  z  =  3?  -  2 

13.  Determine  which  of  the  planes  whose  equations  are 
given  below  are  parallel  and  which  are  perpendicular. 
Are  any  of  the  planes  the  same? 

(a)  2x  +  3y  -  z  =  3 

(b)  -6x  +  4y  -  2z  +  2  =  0 

(c)  x  +  y  -  z  =  2 

(d)  10*  +  I5y  -  5z  =  1 

(e)  3x  -  2y  +  z  =  1 

14.  (a)  What  is  the  angle  between  the  diagonal  of  a  cube 

and  one  of  the  edges  it  meets?  (Hint:  Locate  the 
cube  in  space  in  a  convenient  way.) 

(b)  Find  the  angle  between  the  diagonal  of  a  cube  and 
the  diagonal  of  one  of  its  faces. 

15.  Mark  each  of  the  following  statements  with  a  1  if  you 
agree,  —  1  if  you  disagree: 

(1)  Red  is  my  favorite  color. 

(2)  I  consider  myself  to  be  a  good  athlete. 

(3)  I  like  cats  more  than  dogs. 

(4)  I  enjoy  spicy  foods. 

(5)  Mathematics  is  my  favorite  subject. 

Your  responses  to  the  preceding  "questionnaire"  may 
be  considered  to  form  a  vector  in  R5 .  Suppose  that  you 
and  a  friend  calculate  your  respective  "response  vec- 
tors" for  the  questionnaire.  Explain  the  significance  of 
the  dot  product  of  your  two  vectors. 

16.  The  median  of  a  triangle  is  the  line  segment  that  joins 
a  vertex  of  a  triangle  to  the  midpoint  of  the  opposite 
side.  The  purpose  of  this  problem  is  to  use  vectors  to 
show  that  the  medians  of  a  triangle  all  meet  at  a  point. 

(a)  Using  Figure  1.113,  write  the  vectors  B  M\  and  C  Mi 
in  terms  of  AB  and  AC. 


(b)  Let  P  be  the  point  of  intersection  of  BM\  and  CMi. 
Write  BP  and  CP  in  terms  of  AB  and  AC. 

(c)  Use  the  fact  that  CB  =  CP  +  Jb  =  CA +  AB  to 
show  that  P  must  lie  two-thirds  of  the  way  from  B 
to  Mi  and  two-thirds  of  the  way  from  C  to  Mi. 

(d)  Now  use  part  (c)  to  show  why  all  three  medians  must 
meet  at  P . 

17.  Suppose  that  the  four  vectors  a,  b,  c,  and  d  in  R3  are 
coplanar  (i.e.,  that  they  all  lie  in  the  same  plane).  Show 
that  then  (a  x  b)  x  (c  x  d)  =  0. 

18.  Show  that  the  area  of  the  triangle,  two  of  whose 
sides  are  determined  by  the  vectors  a  and  b  (see 
Figure  1.114),  is  given  by  the  formula 

Area  =  ^7||a||2||b||2  -  (a  •  b)2. 


Figure  1.1 14  The  triangle  in  Exercise  18. 

19.  Let    A(l,3,-1),    5(4,-1,3),    C(2, 5,2),  and 
D(5,  1,  6)  be  the  vertices  of  a  parallelogram. 

(a)  Find  the  area  of  the  parallelogram. 

(b)  Find  the  area  of  the  projection  of  the  parallelogram 
in  the  xy -plane. 

20.  (a)  For  the  line  /  in  R2  given  by  the  equation  ax  + 

by  =  d,  find  a  vector  v  that  is  parallel  to  /. 

(b)  Find  a  vector  n  that  is  normal  to  /  and  has  first 
component  equal  to  a. 

(c)  If  Po(-*o,  Vo)  is  any  point  in  R2,  use  vectors  to  de- 
rive the  following  formula  for  the  distance  from 
P0  to  /: 

\axo  +  bye  —  d\ 

Distance  from  Pq  to  /  = 


Figure  1.113  Two  of  the  three  medians 
of  a  triangle  in  Exercise  16. 


Va2  +  b2 

To  do  this,  you'll  find  it  helpful  to  use  Figure  1.115, 
where  Pi(xu  yi)  is  any  point  on  I. 

(d)  Find  the  distance  between  the  point  (3,  5)  and  the 
line  8jc  —  5y  =  2. 

21.  (a)  If  PoC^o,  yo,  Zo)  is  any  point  in  R3,  use  vectors 
to  derive  the  following  formula  for  the  distance 
from  Po  to  the  plane  n  having  equation  Ax  + 
By  +  Cz  =  D: 

\Ax0  +  By0  +  Czo-  D\ 


Distance  from  Pq  to  fl 


VA2  +  B2  +  C2 


78       Chapter  1  [  Vectors 


y 


Pa 

\  n  / 

I:  ax  +  by  = 

Figure  1.115  Geometric  construction  for 
Exercise  20. 

Figure  1.116  should  help.  (P\{x\,  yi,  Zi)  is  any 


point  in  n.) 

Distance 

~-*P^J 

Yl:Ax  +  By  + 

X 

Figure  1.116  Geometric  construction  for 
Exercise  2 1 . 

(b)  Find  the  distance  between  the  point  (1,5,  —3)  and 
the  plane  x  —  2y  +  2z  +  12  =  0. 

22.  (a)  Let  P  be  a  point  in  space  that  is  not  contained  in  the 

plane  n  that  passes  through  the  three  noncollinear 
points  A ,  B ,  and  C .  Show  that  the  distance  between 
P  and  n  is  given  by  the  expression 

|p-(bxc)| 

lib  x  o || 

where  p  =  AP,  b  =  A§,  and  c  =  AC. 

(b)  Use  the  result  of  part  (a)  to  find  the  distance 
between  (1,  0,  —1)  and  the  plane  containing  the 
points  (1,  2,  3),  (2,  -3,  1),  and  (2,-1,  0). 

23.  Let  A,  B,  C,  and  D  denote  four  distinct  points  in  R3. 

(a)  Show  that  A,  B,  and  C  are  collinear  if  and  only  if 
Xi  x  A~t  =  0. 

(b)  Show  that  A,  B,C,  and  D  are  coplanar  if  and  only 
if  (AT?  x  At)-ci>  =  Q. 

24.  Let  x  =  OP,  the  position  vector  of  a  point  P  in  R3. 
Consider  the  equation 

x-k  1 

II ||  ~  V2' 


Describe  the  configuration  of  points  P  that  satisfy  the 
equation. 

25.  Let  a  and  b  be  two  fixed,  nonzero  vectors  in  R3 ,  and  let 
c  be  a  fixed  constant.  Explain  how  the  pair  of  equations, 

a  •  x  =  c 
a  x  x  =  b , 

completely  determines  the  vector  x  e  R3 . 

26.  (a)  Give  examples  of  vectors  a,  b,  c  in  R3  that  show 

that,  in  general,  it  is  not  true  that  a  x  (b  x  c)  = 
(a  x  b)  x  c.  (That  is,  the  cross  product  is  not  as- 
sociative.) 

(b)  Use  the  Jacobi  identity  (see  Exercise  30  of  §1.4) 
to  show  that,  for  any  vectors  a,  b,  c  in  R3, 

a  x  (b  x  c)  =  (a  x  b)  x  c 

if  and  only  if 

(c  x  a)  x  b  =  0. 

27.  (a)  Given  an  arbitrary  (i.e.,  not  necessarily  regular) 

tetrahedron,  associate  to  each  of  its  four  triangular 
faces  a  vector  outwardly  normal  to  that  face  with 
length  equal  to  the  area  of  that  face.  (See  Fig- 
ure 1.117.)  Show  that  the  sum  of  these  four  vec- 
tors is  zero.  (Hint:  Describe  Vi , . . . ,  V4  in  terms  of 
some  of  the  vectors  that  run  along  the  edges  of  the 
tetrahedron.) 


v3 


Figure  1.117  The  tetrahedron  of  part 
(a)  of  Exercise  27. 

(b)  Recall  that  a  polyhedron  is  a  closed  surface  in 
R3  consisting  of  a  finite  number  of  planar  faces. 
Suppose  you  are  given  the  two  tetrahedra  shown 
in  Figure  1.118  and  that  face  ABC  of  one  is  con- 
gruent to  face  A'  B'C'  of  the  other.  If  you  glue  the 
tetrahedra  together  along  these  congruent  faces, 
then  the  outer  faces  give  you  a  six-faced  polyhe- 
dron. Associate  to  each  face  of  this  polyhedron  an 
outward-pointing  normal  vector  with  length  equal 
to  the  area  of  that  face.  Show  that  the  sum  of  these 
six  vectors  is  zero. 


Miscellaneous  Exercises  for  Chapter  1 


C 


A 


B  B 

Figure  1.118  In  Exercise  27(b),  glue  the  two  tetrahedra  shown  along  congruent  faces. 


28. 


(c)  Outline  a  proof  of  the  following:  Given  an  n-faced 
polyhedron,  associate  to  each  face  an  outward- 
pointing  normal  vector  with  length  equal  to  the 
area  of  that  face.  Show  that  the  sum  of  these  n 
vectors  is  zero. 

Consider  a  right  tetrahedron,  that  is,  a  tetrahedron 
that  has  a  vertex  R  whose  three  adjacent  faces  are  pair- 
wise  perpendicular.  (See  Figure  1.119.)  Use  the  result 
of  Exercise  27  to  show  the  following  three-dimensional 
analogue  of  the  Pythagorean  theorem:  If  a,  b,  and 
c  denote  the  areas  of  the  three  faces  adjacent  to  R 
and  d  denotes  the  area  of  the  face  opposite  R,  then 
d2. 


Figure  1.119  The  right  tetrahedron 
of  Exercise  28.  The  three  faces 
containing  the  vertex  R  are  pairwise 
perpendicular. 

29.  (a)  Use  vectors  to  prove  that  the  sum  of  the  squares 
of  the  lengths  of  the  diagonals  of  a  parallelogram 
equals  the  sum  of  the  squares  of  the  lengths  of  the 
four  sides. 

(b)  Give  an  algebraic  generalization  of  part  (a)  for  R" . 


30.  Show  that  for  any  real  numbers  a\ 
we  have 

n2 


■  ■ ,  an,bi, 

n 


31 .  To  raise  a  square  (n  x  n)  matrix  A  to  a  positive  integer 
power  n,  one  calculates  A"  as  A  ■  A  ■  ■  ■  A  (n  times), 
(a)  Calculate  successive  powers  A,  A2,  A3,  A4  of  the 
1  1 

matrix  A  =     q  \ 


(b)  Conjecture  the  general  form  of  A"  for  the  matrix 
A  of  part  (a),  where  n  is  any  positive  integer. 

(c)  Prove  your  conjecture  in  part  (b)  using  mathemat- 
ical induction. 


32.  A  square  matrix  A  is  called  nilpotent  if  A" 
some  positive  power  n, 

0    1  1 


0  for 


(a)  Show  that  A 


is  nilpotent. 


0  0  0 
0   0  0 

(b)  Use  a  calculator  or  computer  to  show  that  A  = 
0   0   0   0  0 


is  nilpotent. 


^  33. 


The  n  x  n  matrix  Hn  whose  ijth  entry  is  l/(i  +  j  —  1) 
is  called  the  Hilbert  matrix  of  order  n . 

(a)  Write  out  H2,  H],  H^,  H5,  and  H^,  Use  a  com- 
puter to  calculate  their  determinants  exactly.  What 
seems  to  happen  to  det  Hn  as  n  gets  larger? 

(b)  Now  calculate  #10  and  det  H\q.  If  you  use  exact 
arithmetic,  you  should  find  that  det  H\q  /  0  and 
hence  that  //10  is  invertible.  (See  Exercises  30-38 
of  §  1 .6  for  more  about  invertible  matrices.) 

(c)  Now  give  a  numerical  approximation  A  for  H\q. 
Calculate  the  inverse  matrix  B  of  this  approxima- 
tion, if  your  computer  allows.  Then  calculate  AB 
and  B  A .  Do  you  obtain  the  10x10  identity  matrix 
ha  in  both  cases? 

(d)  Explain  what  parts  (b)  and  (c)  suggest  about  the 
difficulties  in  using  numerical  approximations  in 
matrix  arithmetic. 


As  a  child,  you  may  have  played  with  a  popular  toy  called  a 
Spirograph®.  With  it  one  could  draw  some  appealing  geomet- 
ricfigures.  The  Spirograph  consists  of  a  small  toothed  disk  with 
several  holes  in  it  and  a  larger  ring  with  teeth  on  both  inside 
and  outside  as  shown  in  Figure  1.120.  You  can  draw  pictures  by 
meshing  the  small  disk  with  either  the  inside  or  outside  circles 
of  the  ring  and  then  poking  a  pen  through  one  of  the  holes  of 
the  disk  while  turning  the  disk.  (The  large  ring  is  held  fixed.) 


80       Chapter  1  [  Vectors 


An  idealized  version  of  the  Spirograph  can  be  obtained 
by  taking  a  large  circle  (of  radius  a)  and  letting  a  small  circle 
(of  radius  b)  roll  either  inside  or  outside  it  without  slipping. 
A  "Spirograph  "  pattern  is  produced  by  tracking  a  particu- 
lar point  lying  anywhere  on  (or  inside)  the  small  circle.  Exer- 
cises 34-37  concern  this  set-up. 

34.  Suppose  that  the  small  circle  rolls  inside  the  larger 
circle  and  that  the  point  P  we  follow  lies  on  the  circum- 
ference of  the  small  circle.  If  the  initial  configuration 
is  such  that  P  is  at  (a,  0),  find  parametric  equations 
for  the  curve  traced  by  P,  using  angle  t  from  the  posi- 
tive .r-axis  to  the  center  B  of  the  moving  circle.  (This 
configuration  is  shown  in  Figure  1.121.)  The  result- 
ing curve  is  called  a  hypocycloid.  Two  examples  are 
shown  in  Figure  1.122. 


y 


a 

(b.  y\ 

fp ^\ 
At    (  ,  b\ 

Figure  1.121  The  coordinate 
configuration  for  finding  parametric 
equations  for  a  hypocycloid. 


35.  Now  suppose  that  the  small  circle  rolls  on  the  outside 
of  the  larger  circle.  Derive  a  set  of  parametric  equa- 
tions for  the  resulting  curve  in  this  case.  Such  a  curve 
is  called  an  epicycloid,  shown  in  Figure  1.123. 

36.  (a)  A  cusp  (or  corner)  occurs  on  either  the  hypocy- 

cloid or  epicycloid  every  time  the  point  P  on  the 
small  circle  touches  the  large  circle.  Equivalently, 


y 


Figure  1.123  An  epicycloid  with 
a  =  4,b=  1. 


this  happens  whenever  the  smaller  circle  rolls 
through  2tt.  Assuming  that  a/b  is  rational,  how 
many  cusps  does  a  hypocycloid  or  epicycloid 
have?  (Your  answer  should  involve  a  and  b  in  some 
way.) 

(b)  Describe  in  words  and  pictures  what  happens  when 
a/b  is  not  rational. 

37.  Consider  the  original  Spirograph  set-up  again.  If  we 
now  mark  a  point  P  at  a  distance  c  from  the  center 
of  the  smaller  circle,  then  the  curve  traced  by  P  is 
called  a  hypotrochoid  (if  the  smaller  circle  rolls  on 
the  inside  of  the  larger  circle)  or  an  epitrochoid  (if 
the  smaller  circle  rolls  on  the  outside).  Note  that  we 
must  have  b  <  a,  but  we  can  have  c  either  larger  or 
smaller  than  b.  (If  c  <  b,  we  get  a  "true"  Spirograph 
pattern  in  the  sense  that  the  point  P  will  be  on  the 
inside  of  the  smaller  circle.  The  situation  when  c  >  b 
is  like  having  P  mounted  on  the  end  of  an  elongated 
spoke  on  the  smaller  circle.)  Give  a  set  of  parametric 
equations  for  the  curves  that  result  in  this  way.  (See 
Figure  1.124.) 

Exercises  38-43  are  made  feasible  through  the  use  of appropri- 
ate software  for  graphing  in  polar,  cylindrical,  and  spherical 


Miscellaneous  Exercises  for  Chapter  1 


coordinates.  (Note:  When  using  software  for  graphing  in  spher- 
ical coordinates,  be  sure  to  check  the  definitions  that  are  used 
for  the  angles  <p  and  8.) 


x 


Figure  1.124  The  configuration  for  finding 
parametric  equations  for  epitrochoids. 

38.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  =  cos  28. 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  cos  28. 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  cos  2<p. 

(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  cos  28. 

39.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  =  sin2#. 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  sin  2$. 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  sin2<p. 

(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  sin  28.  Compare  the  results  of  this  exercise 
with  those  of  Exercise  38. 

40.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  =  cos  38. 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  cos  38. 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  cos  3<p. 

(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  cos  38. 

41.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  =  sin  38. 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  sin  38. 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  sin  3(p. 


(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  sin  38.  Compare  the  results  of  this  exercise 
with  those  of  Exercise  40. 

42.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  —  1  +  sin  |.  (This  curve  is  known  as  a  nephroid, 
meaning  "kidney  shaped.") 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  1  +  sin  ~ . 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  1  +  sin  | . 

(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  1  +  sin  | . 

43.  (a)  Graph  the  curve  in  R2  whose  polar  equation  is 

r  =  8. 

(b)  Graph  the  surface  in  R3  whose  cylindrical  equation 
is  r  =  8. 

(c)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  <p, 

(d)  Graph  the  surface  in  R3  whose  spherical  equation 
is  p  =  8,  where  jt/2  <  cp  <  n  and  0  <  8  <  4n. 

44.  Consider  the  solid  hemisphere  of  radius  5  pictured  in 
Figure  1.125. 

(a)  Describe  this  solid,  using  spherical  coordinates. 

(b)  Describe  this  solid,  using  cylindrical  coordinates. 


z 


y 


Figure  1.125  The  solid  hemisphere  of 
Exercise  44. 


45.  Consider  the  solid  cylinder  pictured  in  Figure  1.126. 

(a)  Describe  this  solid,  using  cylindrical  coordinates 
(position  the  cylinder  conveniently). 

(b)  Describe  this  solid,  using  spherical  coordinates. 

h  6  - 


Figure  1.126  The  solid 
cylinder  of  Exercise  45. 


2.1  Functions  of  Several 
Variables;  Graphing 
Surfaces 

2.2  Limits 

2.3  The  Derivative 

2.4  Properties;  Higher-order 
Partial  Derivatives 

2.5  The  Chain  Rule 

2.6  Directional  Derivatives  and 
the  Gradient 

2.7  Newton's  Method  (optional) 

True/False  Exercises  for 
Chapter  2 

Miscellaneous  Exercises  for 
Chapter  2 


/ 


X  Y 

Figure  2.1  The  mapping 
nature  of  a  function. 


Differentiation  in 
Several  Variables 


2.1    Functions  of  Several  Variables; 
Graphing  Surfaces 

The  volume  and  surface  area  of  a  sphere  depend  on  its  radius,  the  formulas 
describing  their  relationships  being  V  =  |:rr3  and  S  =  4nr2.  (Here  V  and  S 
are,  respectively,  the  volume  and  surface  area  of  the  sphere  and  r  its  radius.) 
These  equations  define  the  volume  and  surface  area  as  functions  of  the  radius. 
The  essential  characteristic  of  a  function  is  that  the  so-called  independent  variable 
(in  this  case  the  radius)  determines  a  unique  value  of  the  dependent  variable  ( V 
or  S).  No  doubt  you  can  think  of  many  quantities  that  are  determined  uniquely 
not  by  one  variable  (as  the  volume  of  a  sphere  is  determined  by  its  radius)  but 
by  several:  the  area  of  a  rectangle,  the  volume  of  a  cylinder  or  cone,  the  average 
annual  rainfall  in  Cleveland  or  the  national  debt.  Realistic  modeling  of  the  world 
requires  that  we  understand  the  concept  of  a  function  of  more  than  one  variable 
and  how  to  find  meaningful  ways  to  visualize  such  functions. 

Definitions,  Notation,  and  Examples 

A  function,  any  function,  has  three  features:  (1)  a  domain  set  X,  (2)  a  codomain 
set  Y,  and  (3)  a  rule  of  assignment  that  associates  to  each  element  x  in  the 
domain  X  a  unique  element,  usually  denoted  f(x),  in  the  codomain  Y.  We  will 
frequently  use  the  notation  f:X—>  Y  for  a  function.  Such  notation  indicates  all 
the  ingredients  of  a  particular  function,  although  it  does  not  make  the  nature  of 
the  rule  of  assignment  explicit.  This  notation  also  suggests  the  "mapping"  nature 
of  a  function,  indicated  by  Figure  2.1. 

EXAMPLE  1  Abstract  definitions  are  necessary,  but  it  is  just  as  important  that 
you  understand  functions  as  they  actually  occur.  Consider  the  act  of  assigning  to 
each  U.S.  citizen  his  or  her  social  security  number.  This  pairing  defines  a  function: 
Each  citizen  is  assigned  one  social  security  number.  The  domain  is  the  set  of  U.S. 
citizens  and  the  codomain  is  the  set  of  all  nine-digit  strings  of  numbers. 

On  the  other  hand  when  a  university  assigns  students  to  dormitory  rooms,  it 
is  unlikely  that  it  is  creating  a  function  from  the  set  of  available  rooms  to  the  set  of 
students.  This  is  because  some  rooms  may  have  more  than  one  student  assigned 
to  them,  so  that  a  particular  room  does  not  necessarily  determine  a  unique  student 
occupant.  ♦ 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


DEFINITION  1 .1    The  range  of  a  function  f:X  ^  Y  is  the  set  of  those 
elements  of  Y  that  are  actual  values  of  /.  That  is,  the  range  of  /  consists  of 
those  y  in  Y  such  that  y  =  f(x)  for  some  x  in  X. 
Using  set  notation,  we  find  that 

Range  /  =  {y  e  Y  \  y  =  f(x)  for  some  x  e  X] . 


In  the  social  security  function  of  Example  1,  the  range  consists  of  those  nine- 
digit  numbers  actually  used  as  social  security  numbers.  For  example,  the  number 
000-00-0000  is  not  in  the  range,  since  no  one  is  actually  assigned  this  number. 


Figure  2.2  Every  y  e  Y  is  "hit" 
by  at  least  one  x  e  X. 


Figure  2.3  The  element  b  e  Y 
is  not  the  image  of  any  x  £  X. 


DEFINITION  1 .2    A  function  /:  X  ->  Y  is  said  to  be  onto  (or  surjective) 

if  every  element  of  Y  is  the  image  of  some  element  of  X,  that  is,  if  range 
f=Y. 


The  social  security  function  is  not  onto,  since  000-00-0000  is  in  the  codomain 
but  not  in  the  range.  Pictorially,  an  onto  function  is  suggested  by  Figure  2.2.  A 
function  that  is  not  onto  looks  instead  like  Figure  2.3.  You  may  find  it  helpful  to 
think  of  the  codomain  of  a  function  /  as  the  set  of  possible  (or  allowable)  values 
of  /,  and  the  range  of  /  as  the  set  of  actual  values  attained.  Then  an  onto  function 
is  one  whose  possible  and  actual  values  are  the  same. 


DEFINITION  1 .3   A  function  /:  X      Y  is  called  one-one  (or  injective)  if 

no  two  distinct  elements  of  the  domain  have  the  same  image  under  /.  That 
is,  /  is  one-one  if  whenever  x\,  *2  £  X  and  x\  ^  x2,  then  f{x\)  ^  f(xz). 
(See  Figure  2.4.) 


not  one-one 


Figure  2.4  The  figure  on  the  left  depicts  a  one-one  mapping;  the  one 
on  the  right  shows  a  function  that  is  not  one-one. 


One  would  expect  the  social  security  function  to  be  one-one,  but  we  have  heard 
of  cases  of  two  people  being  assigned  the  same  number  so  that,  alas,  apparently 
it  is  not. 

When  you  studied  single-variable  calculus,  the  functions  of  interest  were 
those  whose  domains  and  codomains  were  subsets  of  R  (the  real  numbers).  It 
was  probably  the  case  that  only  the  rule  of  assignment  was  made  explicit;  it  is 
generally  assumed  that  the  domain  is  the  largest  possible  subset  of  R  for  which 
the  function  makes  sense.  The  codomain  is  generally  taken  to  be  all  of  R. 


84       Chapter  2  I  Differentiation  in  Several  Variables 


EXAMPLE  2  Suppose  /:  R  —>  R  is  given  by  /(jc)  =  x2.  Then  the  domain  and 
codomain  are,  explicitly,  all  of  R,  but  the  range  of  /  is  the  interval  [0,  oo).  Thus 
/  is  not  onto,  since  the  codomain  is  strictly  larger  than  the  range.  Note  that  /  is 
not  one-one,  since  f(2)  =  /(— 2)  =  4,  but  2     —2.  ♦ 


EXAMPLE  3  Suppose  g  is  a  function  such  that  g(x)  =  >Jx  —  1.  Then  if  we 
take  the  codomain  to  be  all  of  R,  the  domain  cannot  be  any  larger  than  [1,  oo). 
If  the  domain  included  any  values  less  than  one,  the  radicand  would  be  negative 
and,  hence,  g  would  not  be  real-valued.  ♦ 

Now  we're  ready  to  think  about  functions  of  more  than  one  real  variable.  In 
the  most  general  terms,  these  are  the  functions  whose  domains  are  subsets  X  of 
R"  and  whose  codomains  are  subsets  of  R"',  for  some  positive  integers  n  and  m. 
(For  simplicity  of  notation,  we'll  take  the  codomains  to  be  all  of  R'",  except  when 
specified  otherwise.)  That  is,  such  a  function  is  a  mapping  f:XQ  R"  R™  that 
associates  to  a  vector  (or  point)  x  in  X  a  unique  vector  (point)  f(x)  in  R"! . 


EXAMPLE  4  Let  T:  R3  -*  R  be  defined  by  T(x,  y,  z)  =  xy  +  xz  +  yz.  We 
can  think  of  T  as  a  sort  of  "temperature  function."  Given  a  point  x  =  (x,  y,  z)  in 
R3,  T(x)  calculates  the  temperature  at  that  point.  ♦ 


EXAMPLE  5  Let  L:  R"  ->  R  be  given  by  L(x)  =  ||x|| .  This  is  a  "length  func- 
tion" in  that  it  computes  the  length  of  any  vector  x  in  R".  Note  that  L  is  not 
one-one,  since  L(e,-)  =  L(ey)  =  1,  where  e,  and  e;-  are  any  two  of  the  standard 
basis  vectors  for  R" .  L  also  fails  to  be  onto,  since  the  length  of  a  vector  is  always 
nonnegative.  ♦ 


EXAMPLE  6  Consider  the  function  given  by  N(x)  =  x/||x||  where  x  is  a  vector 
in  R3 .  Note  that  N  is  not  defined  if  x  =  0,  so  the  largest  possible  domain  for  N  is 
R3  —  {0}.  The  range  of  N  consists  of  all  unit  vectors  in  R3.  The  function  N  is  the 
"normalization  function,"  that  is,  the  function  that  takes  a  nonzero  vector  in  R3 
and  returns  the  unit  vector  that  points  in  the  same  direction.  ♦ 


EXAMPLE  7  Sometimes  a  function  may  be  given  numerically  by  a  table. 
One  such  example  is  the  notion  of  windchill — the  apparent  temperature  one 
feels  when  taking  into  account  both  the  actual  air  temperature  and  the  speed  of 
the  wind.  A  standard  table  of  windchill  values  is  shown  in  Figure  2.5. 1  From  it 
we  see  that  if  the  air  temperature  is  20  °F  and  the  windspeed  is  25  mph,  the  wind- 
chill temperature  ("how  cold  it  feels")  is  3  °F.  Similarly,  if  the  air  temperature  is 
35°F  and  the  windspeed  is  10  mph,  then  the  windchill  is  27°F.  In  other  words, 
if  s  denotes  windspeed  and  t  air  temperature,  then  the  windchill  is  a  function 
W(s,  t).  ♦ 

The  functions  described  in  Examples  4,  5,  and  7  are  scalar-valued  functions, 
that  is,  functions  whose  codomains  are  R  or  subsets  of  R.  Scalar- valued  functions 
are  our  main  concern  for  this  chapter.  Nonetheless,  let's  look  at  a  few  examples 
of  functions  whose  codomains  are  R"!  where  m  >  1 . 


NOAA,  National  Weather  Service,  Office  of  Climate,  Water,  and  Weather  Services,  "NWS  Wind  Chill 
Temperature  Index."  February  26,  2004.  <http://www.nws.noaa.gov/om/windchill>  (July  31,  2010). 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


Air  Temp 
(deg  F) 

Windspeed  (mph) 

5 

10 

15 

20 

25 

30 

35  40 

45 

50 

55 

60 

40 

36 

34 

32 

30 

29 

28 

28  27 

26 

26 

25 

25 

35 

31 

27 

25 

24 

23 

22 

21  20 

19 

19 

18 

17 

30 

25 

21 

19 

17 

16 

15 

14  13 

12 

12 

11 

10 

25 

19 

15 

13 

11 

9 

8 

7  6 

5 

4 

4 

3 

20 

13 

9 

6 

4 

3 

1 

0  -1 

-2 

-3 

-3 

-4 

15 

7 

3 

0 

-2 

-4 

-5 

-7  -8 

-9 

-10 

-11 

-11 

10 

1 

-4 

-7 

-9 

-11 

-12 

-14  -15 

-16 

-17 

-18 

-19 

5 

-5 

-10 

-13 

-15 

-17 

-19 

-21  -22 

-23 

-24 

-25 

-26 

0 

-11 

-16 

-19 

-22 

-24 

-26 

-27  -29 

-30 

-31 

-32 

-33 

-5 

-16 

-22 

-26 

-29 

-31 

-33 

-34  -36 

-37 

-38 

-39 

-40 

-10 

-22 

-28 

-32 

-35 

-37 

-39 

-41  -43 

-44 

-45 

-46 

-48 

-15 

-28 

-35 

-39 

-42 

-44 

-46 

-48  -50 

-51 

-52 

-54 

-55 

-20 

-34 

-41 

-45 

-48 

-51 

-53 

-55  -57 

-58 

-60 

-61 

-62 

-25 

-40 

-47 

-51 

-55 

-58 

-60 

-62  -64 

-65 

-67 

-68 

-69 

-30 

-46 

-53 

-58 

-61 

-64 

-67 

-69  -71 

-72 

-74 

-75 

-76 

-35 

-52 

-59 

-64 

-68 

-71 

-73 

-76  -78 

-79 

-81 

-82 

-84 

-40 

-57 

-66 

-71 

-74 

-78 

-80 

-82  -84 

-86 

-88 

-89 

-91 

-45 

-63 

-72 

-77 

-81 

-84 

-87 

-89  -91 

-93 

-95 

-97 

-98 

Figure  2.6  The  helix  of 
Example  8.  The  arrow  shows  the 
direction  of  increasing  t . 


Figure  2.5  Table  of  windchill  values  in  English  units. 

EXAMPLE  8  Define  f:  R  ->  R3  by  f(f)  =  (cos  t,  sin?,  t).  The  range  off  is  the 
curve  in  R3  with  parametric  equations  x  =  cos  t ,  y  =  sin  t ,  z  =  t .  If  we  think  of 
t  as  a  time  parameter,  then  this  function  traces  out  the  corkscrew  curve  (called  a 
helix)  shown  in  Figure  2.6.  ♦ 

EXAMPLE  9  We  can  think  of  the  velocity  of  a  fluid  as  a  vector  in  R3.  This 
vector  depends  on  (at  least)  the  point  at  which  one  measures  the  velocity  and  also 
the  time  at  which  one  makes  the  measurement.  In  other  words,  velocity  may  be 
considered  to  be  a  function  v:  X  c  R4  — >•  R3.  The  domain  X  is  a  subset  of  R4 
because  three  variables  x,  y,  z  are  required  to  describe  a  point  in  the  fluid  and  a 
fourth  variable  t  is  needed  to  keep  track  of  time.  (See  Figure  2.7.)  For  instance, 
such  a  function  v  might  be  given  by  the  expression 


v(x,  y,  z,  t)  =  xyzt\  +  (x1 


y2)l 


(3z  +  /)k. 


Figure  2.7  A  water 
pitcher.  The  velocity  v  of 
the  water  is  a  function 
from  a  subset  of  R4  to  R3. 


You  may  have  noted  that  the  expression  for  v  in  Example  9  is  considerably 
more  complicated  than  those  for  the  functions  given  in  Examples  4-8.  This  is 
because  all  the  variables  and  vector  components  have  been  written  out  explicitly. 
In  general,  if  we  have  a  function  f:  X  c  R"  — >  Rm,  then  xeX  can  be  written  as 
x  =  (xi ,  X2,  ■  ■  ■ ,  x„)  and  f  can  be  written  in  terms  of  its  component  functions 
/i.  fz<  ■  ■  ■ .  fm-  The  component  functions  are  scalar- valued  functions  of  x  £  X 
that  define  the  components  of  the  vector  f(x)  £  R"! .  What  results  is  a  morass  of 
symbols: 

f(x)  =  f(jti,  X2,  ■ .  ■ ,  xn)  (emphasizing  the  variables) 

—  (/i(x)>  /2(x)>  ■  •  •  >  /m(x))       (emphasizing  the  component  functions) 
=  {fi{x\,x2,  . .  .,xn),  f2(xi,x2,  . .  .  ,Xn),        fm(Xl,X2,  ■  ■  .,x„)) 

(writing  out  all  components). 


86       Chapter  2  I  Differentiation  in  Several  Variables 


For  example,  the  function  L  of  Example  5,  when  expanded,  becomes 


L(x)  =  L(x\,x2, . . . ,  x„)  =  ^Jxf  +  x\ 
The  function  N  of  Example  6  becomes 

X  (X],X2,X-i) 

N(x)  =  — 
^  llxll 


x2  x3 


and,  hence,  the  three  component  functions  of  N  are 

X\  X2 

Ni(xi,x2,X3)  =  N2(xi,  x2,  xi)  = 


X^  ~\~  X2  ~\~  X^  J X\  -\-  x2  ~\~  x^ 

x3 

Ni(xi,X2,  x3)  = 


X\  -\-  Xj  ~"T~  X-i 


Although  writing  a  function  in  terms  of  all  its  variables  and  components  has 
the  advantage  of  being  explicit,  quite  a  lot  of  paper  and  ink  are  used  in  the  process. 
The  use  of  vector  notation  not  only  saves  space  and  trees  but  also  helps  to  make 
the  meaning  of  a  function  clear  by  emphasizing  that  a  function  maps  points  in 
R"  to  points  in  R'" .  Vector  notation  makes  a  function  of  300  variables  look  "just 
like"  a  function  of  one  variable.  Try  to  avoid  writing  out  components  as  much  as 
you  can  (except  when  you  want  to  impress  your  friends). 


Visualizing  Functions   

No  doubt  you  have  been  graphing  scalar-valued  functions  of  one  variable  for  so 
long  that  you  give  the  matter  little  thought.  Let's  scrutinize  what  you've  been  do- 
ing, however.  A  function  /:XcR^  R  takes  a  real  number  and  returns  another 
real  number  as  suggested  by  Figure  2.8.  The  graph  of  /  is  something  that  "lives" 
in  R2.  (See  Figure  2.9.)  It  consists  of  points  (x,  y)  such  that  y  =  f(x).  That  is, 

Graph  /  =  {{x,  f(x))  \  x  e  X}  =  {(x,  y)\xeX,y  =  f(x)} . 

The  important  fact  is  that,  in  general,  the  graph  of  a  scalar-valued  function 
of  a  single  variable  is  a  curve — a  one-dimensional  object — sitting  inside  two- 
dimensional  space. 


y 


f 


SKA 

(x,f(x))  / 

X 

Figure  2.8  A  function  /:ICR->R,  Figure  2.9  The  graph  off. 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


Now  suppose  we  have  a  function  /:XcR2->R,  that  is,  a  function  of  two 
variables.  We  make  essentially  the  same  definition  for  the  graph: 


Graph  /  =  {(x,  /(x))  xel).  (1) 


Of  course,  x  =  (x,  y)  is  a  point  of  R2.  Thus,  {(x,  fix))}  may  also  be  written  as 

l(x,  y,  fix,  y))} ,    or  as    {(x,  y,  z)  |  (x,  y)  e  X,  z  =  f(x,  y)} . 

Hence,  the  graph  of  a  scalar-valued  function  of  two  variables  is  something  that 
sits  in  R3.  Generally  speaking,  the  graph  will  be  a  surface. 

EXAMPLE  1 0   The  graph  of  the  function 

9  1,1,7 
/:R2^R      /(x,y)=-y3-y--x2  +  - 

is  shown  in  Figure  2.10.  For  each  point  x  =  (x,  y)  in  R2,  the  point  in  R3  with 
coordinates  (x,  y,  ^y3  —  y  —  \x2  +  |)  is  graphed.  ♦ 


Figure  2.10  The  graph  of  fix,  y)  =  -j^y3  —  y  —  \x2  +  |. 


Graphing  functions  of  two  variables  is  a  much  more  difficult  task  than  graph- 
ing functions  of  one  variable.  Of  course,  one  method  is  to  let  a  computer  do 
the  work.  Nonetheless,  if  you  want  to  get  a  feeling  for  functions  of  more  than 
one  variable,  being  able  to  sketch  a  rough  graph  by  hand  is  still  a  valuable  skill. 
The  trick  to  putting  together  a  reasonable  graph  is  to  find  a  way  to  cut  down  on 
the  dimensions  involved.  One  way  this  can  be  achieved  is  by  drawing  certain 
special  curves  that  lie  on  the  surface  z  =  fix,  y).  These  special  curves,  called 
contour  curves,  are  the  ones  obtained  by  intersecting  the  surface  with  horizontal 
planes  z  =  c  for  various  values  of  the  constant  c.  Some  contour  curves  drawn 
on  the  surface  of  Example  10  are  shown  in  Figure  2.1 1.  If  we  compress  all  the 
contour  curves  onto  the  xy-plane  (in  essence,  if  we  look  down  along  the  posi- 
tive z-axis),  then  we  create  a  "topographic  map"  of  the  surface  that  is  shown  in 
Figure  2.12.  These  curves  in  the  xy-plane  are  called  the  level  curves  of  the  original 
function  /. 

The  point  of  the  preceding  discussion  is  that  we  can  reverse  the  process  in 
order  to  sketch  systematically  the  graph  of  a  function  /  of  two  variables:  We 


88       Chapter  2  I  Differentiation  in  Several  Variables 


Figure  2.1 1  Some  contour  curves  of  the  Figure  2.1 2  Some  level  curves  of 

function  in  Example  10.  the  function  in  Example  10. 


first  construct  a  topographic  map  in  R2  by  finding  the  level  curves  of  /,  then 
situate  these  curves  in  R3  as  contour  curves  at  the  appropriate  heights,  and  finally 
complete  the  graph  of  the  function.  Before  we  give  an  example,  let's  restate  our 
terminology  with  greater  precision. 


DEFINITION  1 .4  Let  /:XcR2^Rbea  scalar-valued  function  of  two 
variables.  The  level  curve  at  height  c  of  /  is  the  curve  in  R2  defined  by  the 
equation  f(x,  y)  =  c,  where  c  is  a  constant.  In  mathematical  notation, 

Level  curve  at  height  c  =  {(x,  y)  £  R2  |  fix,  y)  =  c]  . 

The  contour  curve  at  height  c  of  /  is  the  curve  in  R3  defined  by  the  two 
equations  z  =  fix,  y)  and  z  =  c.  Symbolized, 

Contour  curve  at  height  c  =  {(x,  y,  z)  £  R3  |  z  =  fix,  y)  =  c}  . 

In  addition  to  level  and  contour  curves,  consideration  of  the  sections  of  a 
surface  by  the  planes  where  x  or  y  is  held  constant  is  also  helpful.  A  section  of  a 
surface  by  a  plane  is  just  the  intersection  of  the  surface  with  that  plane.  Formally, 
we  have  the  following  definition: 


DEFINITION  1 .5  Let  /:  X  C  R2  ->  R  be  a  scalar-valued  function  of  two 
variables.  The  section  of  the  graph  of  /  by  the  plane  x  =  c  (where  c 
is  a  constant)  is  the  set  of  points  (x,  y,  z),  where  z  =  f(x,  y)  and  x  =  c. 
Symbolized, 

Section  by  x  =  c  is  {(x,  y,  z)  e  R3  I  z  =  f(x,  y),  x  =  c}. 

Similarly,  the  section  of  the  graph  of  /  by  the  plane  y  =  c  is  the  set  of 

points  described  as  follows: 

Section  by  y  =  c  is  {(x,  y,  z)  e  R3  |  z  =  fix,  y),  y  =  c]. 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


EXAMPLE  1 1  We'll  use  level  and  contour  curves  to  construct  the  graph  of  the 
function 

/:R2^R,      f(x,y)  =  4-x2-y2. 

By  Definition  1 .4,  the  level  curve  at  height  c  is 

{(x,  y)  6  R2  |  4  -  x2  -  y2  =  c}  =  {(x,  y)  |  x2  +  y2  =  4  -  c} . 

Thus,  we  see  that  the  level  curves  for  c  <  4  are  circles  centered  at  the  origin  of 
radius  -J4  —  c.  The  level  "curve"  at  height  c  =  4  is  not  a  curve  at  all  but  just  a 
single  point  (the  origin).  Finally,  there  are  no  level  curves  at  heights  larger  than  4 
since  the  equation  x2  +  y2  =  4  —  c  has  no  real  solutions  in  x  and  y.  (Why  not?) 
These  remarks  are  summarized  in  the  following  table: 


c 

Level  curve  x2  +  y2  =  4  —  c 

-5 

x2  +  y2  =  9 

-1 

x2  +  y2  =  5 

0 

x2  +  y2  =  4 

1 

x2  +  y2  =  3 

3 

x2  +  y2  =  l 

4 

x2  +  y2  =  0               x  =  y  =  0 

c,  where  c  >  4 

empty 

Thus,  the  family  of  level  curves,  the  "topographic  map"  of  the  surface  z  = 
4  —  x2  —  y2,  is  shown  in  Figure  2.13.  Some  contour  curves,  which  sit  in  R3, 
are  shown  in  Figure  2.14,  where  we  can  get  a  feeling  for  the  complete  graph  of 
z  =  4  —  x2  —  y2.  It  is  a  surface  that  looks  like  an  inverted  dish  and  is  called  a 
paraboloid.  (See  Figure  2. 15.)  To  make  the  picture  clearer,  we  have  also  sketched 
in  the  sections  of  the  surface  by  the  planes  x  =  0  and  y  =  0.  The  section  by  x  =  0 
is  given  analytically  by  the  set 

{(x,  y,  z)  e  R3  |  z  =  4  -  x2  -  y2,  x  =  0}  =  {(0,  y,z)\z  =  4-  y2}  . 

Similarly,  the  section  by  y  =  0  is 

{(x,  y,  z)  e  R3  |  z  =  4  -  x2  -  y2,  y  =  0}  =  {(x,  0,  z)  \  z  =  4  -  x2}  . 


Figure  2.13  The  topographic  Figure  2.14  Some  contour  Figure  2.15  The  graph  of 

map  of  z  =  4  —  x2  —  y2  (i.e.,  curves  of  z  =  4  —  x2  —  y2.  f(x,  y)  =  4  —  x2  —  y2 . 

several  of  its  level  curves). 


Chapter  2  |  Differentiation  in  Several  Variables 


Since  these  sections  are  parabolas,  it  is  easy  to  see  how  this  surface  obtained  its 
name.  ♦ 

EXAMPLE  12  We'll  graph  the  function  g:  R2  ->  R,  g(x,  y)  =  y2  -  x2.  The 
level  curves  are  all  hyperbolas,  with  the  exception  of  the  level  curve  at  height  0, 
which  is  a  pair  of  intersecting  lines. 


c  =  -4 


c 

Level  curve  y2  —  x2  =  c 

-4 

x2  -  y2  =  4 

4 

-1 

x2  -  y2  =  1 

0 

y2-X2=0 

<S=>     (y-x)(y+x)  =  0  J= 

=>■    y  =  ±x 

X 

1 

y2-x2  =  \ 

4 

y2-X2=4 

Figure  2.1 6  Some  level  curves 
of  g(x,y)  =  y2  -x2. 


Figure  2.1 7  The  contour  curves 
and  graph  of  g(x,  y)  =  y2  —  x2. 


The  collection  of  level  curves  is  graphed  in  Figure  2.16.  The  sections  by  x  =  c 
are 

l(x,  y,  z)  |  z  =  y2  -  x2,x  =  c}  =  {(c,  y,  z)  \  z  =  y2  -  c2}. 

These  are  clearly  parabolas  in  the  planes  x  =  c.  The  sections  by  y  =  c  are 

{(x,  y,  z)  |  z  =  y2  -  x2,  y  =  c)  =  {(c,  y,  z)  |  z  =  c2  -  x2}, 

which  are  again  parabolas.  The  level  curves  and  sections  generate  the  contour 
curves  and  surface  depicted  in  Figure  2.17.  Perhaps  understandably,  this  surface 
is  called  a  hyperbolic  paraboloid.  ♦ 

EXAMPLE  13  We  compare  the  graphs  of  the  function  f(x,  y)  =  4  —  x2  —  y2 
of  Example  1 1  with  that  of 

h:  R2  -  {(0,  0)}  ->  R,      h(x,  y)  =  ln(x2  +  y2). 

The  level  curve  of  h  at  height  c  is 

{(jc,  y)  6  R2  |  ln(x2  +  y2)  =  c]  =  {(x,  y)  |  x2  +  y2  =  ec\  . 

Since  ec  >  0  for  all  c  6  R,  we  see  that  the  level  curve  exists  for  any  c  and  is  a 
circle  of  radius  J~e^  =  ec/2. 


c 

Level  curve  x2  +  y2  =  ec 

-5 

x2  +  y2  =  e~5 

-1 

x2  +  y2  =  e~l 

0 

x 2  +  y2  =  1 

1 

x2  +  y2  =  e 

3 

x2  +  y2  =  e3 

4 

x2+y2  =  e4 

The  collection  of  level  curves  is  shown  in  Figure  2.18  and  the  graph  in  Figure  2.19. 
Note  that  the  section  of  the  graph  by  x  =  0  is 

{(x,y,z)  eR3  |  Z  =  \n(x2  +  y2),x  =  0}  =  {(0,  y,  z)  \  z  =  ln(y2)  =  2 In  | y | }  . 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


y 


X 


Figure  2.1 8  The  collection  of  level  curves  ofz  =  ln(x2  + v2). 


The  section  by  _y  =  0  is  entirely  similar: 

{(x,y,z)eR3  |  z  =  ln(x2  +  y2),  y  =  0}  =  {(*,(>,  z)|  z  =  ln(x2)  =  2 In  \x\ }  . 

♦ 

In  fact,  if  we  switch  from  Cartesian  to  cylindrical  coordinates,  it  is  quite 
easy  to  understand  the  surfaces  in  both  Examples  11  and  13.  In  view  of  the 
Cartesian/cylindrical  relation  x2  +  y2  =  r2,  we  see  that  for  the  function  /  of 
Example  11, 

z  =  4  -  x2  -  y2  =  4  -  (x2  +  y2)  =  4  -  r2. 

For  the  function  h  of  Example  13,  we  have 

z  =  ln(x2  +  y2)  =  ln(r2)  =  21nr, 

where  we  assume  the  usual  convention  that  the  cylindrical  coordinate  r  is  non- 
negative.  Thus  both  of  the  graphs  in  Figures  2.15  and  2.19  are  of  surfaces  of 
revolution  obtained  by  revolving  different  curves  about  the  z-axis.  As  a  result, 
the  level  curves  are,  in  general,  circular. 

The  preceding  discussion  has  been  devoted  entirely  to  graphing  scalar- valued 
functions  of  just  two  variables.  However,  all  the  ideas  can  be  extended  to  more 
variables  and  higher  dimensions.  If  /:  X  c  R"  —>  R  is  a  (scalar-valued)  function 
of  n  variables,  then  the  graph  of  /  is  the  subset  of  R"+1  given  by 


Graph  /  =  {(x,  /(x))  xel} 

=  {{x\,  ■  ■  ■  ,  X„,  Xn+\) 

(Xi,  .  .  .,Xn)  €  X, 

(2) 

xn+\  =  f(xl ,  ■  ■  ■  ,  Xn)}  . 

92       Chapter  2  |  Differentiation  in  Several  Variables 


z 


y 


Figure  2.20  The  level  sets  of  the 
function  F(x,  y,z)  =  x  +  y  +  z 
are  planes  in  R3 . 


y 


(1,0) 

i  x 


Figure  2.21  The  unit  circle 
x2  +  y2  =  1. 


Z 


Figure  2.22  The  sphere  of  radius 
a,  centered  at  (xq,  yo,  zo). 


(The  compactness  of  vector  notation  makes  the  definition  of  the  graph  of  a  function 
of  n  variables  exactly  the  same  as  in  ( 1 ).)  The  level  set  at  height  c  of  such  a  function 
is  defined  by 

Level  set  at  height  c  =  {x  e  R"  |  f(x)  =  c} 

=  {(*i,  X2,  .  .  .  ,  X„)  |  f(Xl,  X2,  ...,X»)  =  C}. 

While  the  graph  of  /  is  a  subset  of  R"+1,  a  level  set  of  /  is  a  subset  of  R".  This 
makes  it  possible  to  get  some  geometric  insight  into  graphs  of  functions  of  three 
variables,  even  though  we  cannot  actually  visualize  them. 

EXAMPLE  14  Let  F:  R3  ->  R  be  given  by  F(x,  y,  z)  =  x  +  y  +  z.  Then  the 
graph  of  F  is  the  set  {(*,  y,  z,  if)  |  w  =  x  +  y  +  z}  and  is  a  subset  (called  a 
hypersurface)  of  R4,  which  we  cannot  depict  adequately.  Nonetheless,  we  can 
look  at  the  level  sets  of  F,  which  are  surfaces  in  R3.  (See  Figure  2.20.)  We  have 

Level  set  at  height  c  =  {(x,  y,  z)  \  x  +  y  +  z  =  c}. 

Thus,  the  level  sets  form  a  family  of  parallel  planes  with  normal  vector  i  +  j  +  k. 

♦ 

Surfaces  in  General  

Not  all  curves  in  R2  can  be  described  as  the  graph  of  a  single  function  of  one 
variable.  Perhaps  the  most  familiar  example  is  the  unit  circle  shown  in  Figure  2.2 1 . 
Its  graph  cannot  be  determined  by  a  single  equation  of  the  form  y  =  f(x)  (or,  for 
that  matter,  by  one  of  the  form  x  =  g(y)).  As  we  know,  the  graph  of  the  circle  may 
be  described  analytically  by  the  equation  x2  +  y2  =  1.  In  general,  a  curve  in  R2 
is  determined  by  an  arbitrary  equation  in  x  and  y,  not  necessarily  one  that  isolates 
y  alone  on  one  side.  In  other  words,  this  means  that  a  general  curve  is  given  by  an 
equation  of  the  form  F(x,  y)  —  c  (i.e.,  a  level  set  of  a  function  of  two  variables). 

The  analogous  situation  occurs  with  surfaces  in  R3 .  Frequently  a  surface  is 
determined  by  an  equation  of  the  form  F(x,  y,  z)  —  c  (i.e.,  as  a  level  set  of  a 
function  of  three  variables),  not  necessarily  one  of  the  form  z  =  f(x,  y). 

EXAMPLE  15  A  sphere  is  a  surface  in  R3  whose  points  are  all  equidistant  from 
a  fixed  point.  If  this  fixed  point  is  the  origin,  then  the  equation  for  the  sphere  is 

llx-OH  =  |WI  =o,  (3) 

where  a  is  a  positive  constant  and  x  =  (x,  y,  z)  is  a  point  on  the  sphere.  If  we 
square  both  sides  of  equation  (3)  and  expand  the  (implicit)  dot  product,  then  we 
obtain  perhaps  the  familiar  equation  of  a  sphere  of  radius  a  centered  at  the  origin: 

x2  +  y2  +  Z2  =  a2.  (4) 

If  the  center  of  the  sphere  is  at  the  point  xo  =  (jto,  yo,  zq),  rather  than  the  origin, 
then  equation  (3)  should  be  modified  to 

l|x-x0||=a.  (5) 

(See  Figure  2.22.) 

When  equation  (5)  is  expanded,  the  following  general  equation  for  a  sphere 
is  obtained: 


(x  -  x0)2  +  (y-  y0)2  +  (z  -  zo)2  =  a2  (6) 


2.1  |  Functions  of  Several  Variables;  Graphing  Surfaces 


In  the  equation  for  a  sphere,  there  is  no  way  to  solve  for  z  uniquely  in  terms 
of  x  and  y.  Indeed,  if  we  try  to  isolate  z  in  equation  (4),  then 

2         2         2  2 

z  =  a  —  x  —  y  , 

so  we  are  forced  to  make  a  choice  of  positive  or  negative  square  roots  in  order  to 
solve  for  z: 

z  =  \J  a1  —  x2  —  y2     or     z.  =  —\J  a2  —  x2  —  y2. 

The  positive  square  root  corresponds  to  the  upper  hemisphere  and  the  negative 
square  root  to  the  lower  one.  In  any  case,  the  entire  sphere  cannot  be  the  graph 
of  a  single  function  of  two  variables.  ♦ 

Of  course,  the  graph  of  a  function  of  two  variables  does  describe  a  surface  in 
the  "level  set"  sense.  If  a  surface  happens  to  be  given  by  an  equation  of  the  form 

z  =  f(x,  y) 

for  some  appropriate  function  /:XcR2^R,  then  we  can  move  z  to  the  oppo- 
site side,  obtaining 

f{x,  y)-z  =  0. 
If  we  define  a  new  function  F  of  three  variables  by 

F{x,y,z)  =  f(x,y)-z, 

then  the  graph  of  /  is  precisely  the  level  set  at  height  0  of  F.  We  reiterate  this 
point  since  it  is  all  too  often  forgotten:  The  graph  of  a  function  of  two  variables  is 
a  surface  in  R3  and  is  a  level  set  of  a  function  of  three  variables.  However,  not  all 
level  sets  of functions  of  three  variables  are  graphs  of functions  of  two  variables. 
We  urge  you  to  understand  this  distinction. 

Quadric  Surfaces 

Conic  sections,  those  curves  obtained  from  the  intersection  of  a  cone  with  various 
planes,  are  among  the  simplest,  yet  also  the  most  interesting,  of  plane  curves: 
They  are  the  circle,  the  ellipse,  the  parabola,  and  the  hyperbola.  Besides  being 
produced  in  a  similar  geometric  manner,  conic  sections  have  an  elegant  algebraic 
connection:  Every  conic  section  is  described  analytically  by  a  polynomial  equation 
of  degree  two  in  two  variables.  That  is,  every  conic  can  be  described  by  an  equation 
that  looks  like 

Ax2  +  Bxy  +  Cy2  +  Dx  +  Ey  +  F  =  0 

for  suitable  constants  A, . . . ,  F. 

In  R3,  the  analytic  analogue  of  the  conic  section  is  called  a  quadric  surface. 
Quadric  surfaces  are  those  defined  by  equations  that  are  polynomials  of  degree 
two  in  three  variables: 

Ax2  +  Bxy  +  Cxz  +  Dy2  +  Eyz  +  Fz2  +  Gx  +  Hy  +  lz  +  J  =  0. 

To  pass  from  this  equation  to  the  appropriate  graph  is,  in  general,  a  cumbersome 
process  without  the  aid  of  either  a  computer  or  more  linear  algebra  than  we 
currently  have  at  our  disposal.  So,  instead  we  offer  examples  of  those  quadric 
surfaces  whose  axes  of  symmetry  lie  along  the  coordinate  axes  in  R3  and  whose 
corresponding  analytic  equations  are  relatively  simple.  In  the  discussion  that 
follows,  a,  b,  and  c  are  constants,  which,  for  convenience,  we  take  to  be  positive. 


94       Chapter  2  |  Differentiation  in  Several  Variables 


Z 


Figure  2.26  The  elliptic 


cone 


Figure  2.27  The  graph  of  the 

2  2  2 

x      y  z 
equation  — r  +  -pr  T  =  1  is  a 

a2-      bL  cL 

hyperboloid  of  one  sheet. 


Ellipsoid  (Figure  2.23.)  x2/a2  +  y2/b2  +  z2/c2  =  1. 
This  is  the  three-dimensional  analogue  of  an  ellipse  in  the  plane.  The  sections 
of  the  ellipsoid  by  planes  perpendicular  to  the  coordinate  axes  are  all  ellipses. 
For  example,  if  the  ellipsoid  is  intersected  with  the  plane  z  =  0,  one  obtains  the 
standard  ellipse  x2/a2  +  y2/b2  =  1,  z  =  0.  If  a  =  b  =  c,  then  the  ellipsoid  is  a 
sphere  of  radius  a. 

Elliptic  paraboloid  (Figure  2.24.)    z/c  =  x2/a2  +  y2/b2. 

(The  roles  of  x,  y,  and  z  may  be  interchanged.)  This  surface  is  the  graph  of 
a  function  of  x  and  y.  The  paraboloid  has  elliptical  (or  single-point  or  empty) 
sections  by  the  planes  "z  =  constant"  and  parabolic  sections  by  "x  =  constant"  or 
"  v  =  constant"  planes.  The  constants  a  and  b  affect  the  aspect  ratio  of  the  elliptical 
cross  sections,  and  the  constant  c  affects  the  steepness  of  the  dish.  (Larger  values 
of  c  produce  steeper  paraboloids.) 

Hyperbolic  paraboloid  (Figure  2.25.)    z/c  =  y2/b2  —  x2/a2. 

(Again  the  roles  of  x,  y,  and  z  may  be  interchanged.)  We  saw  the  graph  of  this 
surface  earlier  in  Example  12  of  this  section.  It  is  shaped  like  a  saddle  whose  "x  = 
constant"  or  "y  =  constant"  sections  are  parabolas  and  "z  =  constant"  sections 
are  hyperbolas. 

Elliptic  cone  (Figure  2.26.)    z2/c2  =  x2/a2  +  y2/b2. 

The  sections  by  "z  =  constant"  planes  are  ellipses.  The  sections  by  x  =  0  or 

y  =  0  are  each  a  pair  of  intersecting  lines. 

Hyperboloid  of  one  sheet  (Figure  2.27.)    x2/a2  +  y2/b2  -  z2/c2  =  1. 

The  term  "one  sheet"  signifies  that  the  surface  is  connected  (i.e.,  that  you  can 
travel  between  any  two  points  on  the  surface  without  having  to  leave  the  surface). 
The  sections  by  "z  =  constant"  planes  are  ellipses  and  those  by  "x  =  constant" 
or  "y  =  constant"  planes  are  hyperbolas,  hence,  this  surface's  name. 

Hyperboloid  of  two  sheets  (Figure  2.28.)    z2/c2  -  x2/a2  -  y2/b2  =  1. 

The  fact  that  the  left-hand  side  of  the  defining  equation  is  the  opposite  of  the  left 
side  of  the  equation  for  the  previous  hyperboloid  is  what  causes  this  surface  to 
consist  of  two  pieces  instead  of  one.  More  precisely,  consider  the  sections  of  the 


2.1  |  Exercises  95 


(0,  0,  c) 


Figure  2.28  The  graph  of 
the  equation 

2  2  2 

z     x  y 

cL      or  bL 
hyperboloid  of  two  sheets. 


Figure  2.29  The 

hyperboloids 

asymptotic  to  the  cone 
c2~  a2  +  b2' 


surface  by  planes  of  the  form  z  =  k  for  different  constants  k.  These  sections  are 
thus  given  by 


or,  equivalently,  by 


7i 


2  n 
X  V 

 I-  — 

a2  b2 


r 
b2 

e 


=  1, 


z  =  k 


1,    z  =  k. 


If  —c  <  k  <  c,  then  0  <  k2/c2  <  1.  Thus,  /:2/c2  —  1  <  0,  and  so  the  preceding 
equation  has  no  solution  in  x  and  y.  Hence,  the  section  by  z  =  k,  where  \k\  <  c, 
is  empty.  If  \k\  >  c,  then  the  section  is  an  ellipse.  The  sections  by  "x  =  constant" 
or  "y  =  constant"  planes  are  hyperbolas. 


In  the  same  way  that  the  hyperbolas 

2  n 

x  y 
a2  b2 

are  asymptotic  to  the  lines  y  =  ±{b/a)x,  the  hyperboloids 


=  ±1 


2  2  1 

Z  _  x  yL 
c2      a2  b2 


±  1 


are  asymptotic  to  the  cone 


=  X-,  +  y 


b2' 


This  is  perhaps  intuitively  clear  from  Figure  2.29,  but  let's  see  how  to  prove  it 
rigorously.  In  our  present  context,  to  say  that  the  hyperboloids  are  asymptotic 
to  the  cones  means  that  they  look  more  and  more  like  the  cones  as  |z|  becomes 
(arbitrarily)  large.  Analytically,  this  should  mean  that  the  equations  for  the  hy- 
perboloids should  approximate  the  equation  for  the  cone  for  sufficiently  large  \z\. 
The  equations  of  the  hyperboloids  can  be  written  as  follows: 

r2        v2        72  72  /  r1 

x  .  y    *  ±1  *  (i±* 


a2     b2      cA  c-  \  zA 

As  \z\  ->  oo,  c2/z2  ->  0,  so  the  right  side  of  the  equation  for  the  hyperboloids 
approaches  z2/c2.  Hence,  the  equations  for  the  hyperboloids  approximate  that  of 
the  cone,  as  desired. 


2.1  Exercises 


1.  Let/:R^  R  be  given  by  f(x)  =  2x2  +  1. 

(a)  Find  the  domain  and  range  of  /. 

(b)  Is  /  one-one? 

(c)  Is  /  onto? 

2.  Let  g:  R2  ->  Rbe  given  by  g(x,  y)  =  2x2  +  3y2  —  7. 

(a)  Find  the  domain  and  range  of  g. 

(b)  Find  a  way  to  restrict  the  domain  to  make  a  new 
function  with  the  same  rule  of  assignment  as  g  that 
is  one-one. 


(c)  Find  a  way  to  restrict  the  codomain  to  make  a  new 
function  with  the  same  rule  of  assignment  as  g  that 
is  onto. 

Find  the  domain  and  range  of  each  of  the  functions  given  in 
Exercises  3-7. 

3.  fix,  y)=- 

y 

4.  f(x,  y)  =  ln(x  +  y) 

5.  g(x,  y,  z)  =  Jx2  +  (y-  2)2  +  (z  +  I)2 


Chapter  2  |  Differentiation  in  Several  Variables 


6.  g(x,  y,  z)  = 


7.  f(x,y)=  [x  +  y 


y-1 


x2+y2 


8.  Let  f:R2  ->  R3  be  defined  by  f(x,  y)  =  (x  +  y, 
yex,  x2y  +  7).  Determine  the  component  functions 
off. 

9.  Determine  the  component  functions  of  the  function  v 
in  Example  9. 

1 0.  Let  f :  R3  ->  R3  be  defined  by  f(x)  =  x  +  3  j .  Write  out 
the  component  functions  of  f  in  terms  of  the  compo- 
nents of  the  vector  x. 

1 1 .  Consider  the  mapping  that  assigns  to  a  nonzero  vector 
x  in  R3  the  vector  of  length  2  that  points  in  the  direction 
opposite  to  x. 

(a)  Give  an  analytic  (symbolic)  description  of  this 
mapping. 

(b)  If  x  =  (x,  y,  z),  determine  the  component  func- 
tions of  this  mapping. 

1 2.  Consider  the  function  f:  R2  —>  R3  given  by  f(x)  =  Ax, 
2    -1  " 

where  A  =       5  0 
-6  3 


and  the  vector  x  in  R  is 


written  as  the  2x1  column  matrix  x 


X] 
X2 


(a)  Explicitly  determine  the  component  functions  of 
f  in  terms  of  the  components  X\,  X2  of  the  vector 
(i.e.,  column  matrix)  x. 

(b)  Describe  the  range  of  f. 

1 3.  Consider  the  function  f:  R4  — >  R3  given  by  f(x)  =  Ax, 
~  2   0   -1  1 
0   3      0  0 
2  0-11 


and  the  vector  x  in  R4 


is  written  as  the  4x1  column  matrix  x 


x\ 

Xl 
-V3 
X4 


(a)  Determine  the  component  functions  of  f  in  terms 
of  the  components  xi ,  xi,  *3,  X4  of  the  vector  (i.e., 
column  matrix)  x. 

(b)  Describe  the  range  of  f. 

In  each  of  Exercises  14-23,  (a)  determine  several  level  curves 
of  the  given  function  f  (make  sure  to  indicate  the  height  c  of 
each  curve);  (b)  use  the  information  obtained  in  part  (a)  to 
sketch  the  graph  of  f. 


14.  f(x,  y)  =  3 

15.  f(x,y) 

16.  f(x,y)  =  x2  +  y2 


x2  +  v2 


17. 

f(x,y)  = 

18. 

fix  y)  = 

Ax2 

19. 

fix,  y)  = 

xy 

20. 

fix,  y)  = 

y 

X 

21. 

fix,  y)  = 

X 

y 

22. 

fix,  y)  = 

3  - 

23. 

fix,  y)  = 

\x\ 

In  Exercises  24-27,  use  a  computer  to  provide  a  portrait  of 
the  given  function  g(x,  y).  To  do  this,  (a)  use  the  computer  to 
help  you  understand  some  of  the  level  curves  of  the  function, 
and  (b)  use  the  computer  to  graph  (a  portion  of)  the  surface 
Z  =  g(x,  y).  In  addition,  mark  on  your  surface  some  of  the 
contour  curves  corresponding  to  the  level  curves  you  obtained 
in  part  (a).  (See  Figures  2.10  and  2.11.) 

^  24.  g(x,  y)  =  yex 

^  25.  g(x,y)  =  x2  -  xy 

^  26.  g(x,  y)  =  ix2  +  3y2y-x2-vl 

^  sin(2  —  x2  —  v2) 

O  27.  g{x,  y)  =      \  . 

x £  +  yL  +  1 

28.  The  ideal  gas  law  is  the  equation  PV  =  kT,  where  P 
denotes  the  pressure  of  the  gas,  V  the  volume,  T  the 
temperature,  and  k  is  a  positive  constant. 

(a)  Describe  the  temperature  T  of  the  gas  as  a  function 
of  volume  and  pressure.  Sketch  some  level  curves 
for  this  function. 

(b)  Describe  the  volume  V  of  the  gas  as  a  function 
of  pressure  and  temperature.  Sketch  some  level 
curves. 

29.  (a)  Graph  the  surfaces  z  =  x2  and  z  =  y2 ■ 

(b)  Explain  how  one  can  understand  the  graph  of  the 
surfaces  z  =  fix)  and  z  =  /(y)  by  considering 
the  curve  in  the  ifii-plane  given  by  v  =  /(«). 

(c)  Graph  the  surface  in  R3  with  equation  y  =  x2. 

30.  Use  a  computer  to  graph  the  family  of  level  curves  for 
the  functions  in  Exercises  20  and  21  and  compare  your 
results  with  those  obtained  by  hand  sketching.  How  do 
you  account  for  any  differences? 

31 .  Given  a  function  f(x ,  y),  can  two  different  level  curves 
of  /  intersect?  Why  or  why  not? 

In  Exercises  32—36,  describe  the  graph  of  g(x,  y,  z)  by  com- 
puting some  level  surfaces.  (If  you  prefer,  use  a  computer  to 
assist  you.) 

32.  g(x,  y,z)  =  x  —  2y  +  3z 

33.  g(x,  y,z)  =  x2  +  y2  —  z 


2.2  |  Limits 


34.  g(x,  y,z)  =  x2  +  y2  +  z2 

35.  g(x,  y,z)  =  x2  +  9y2  +  4z2 

36.  g(x,  y,  z)  =  xy  -  yz 

37.  (a)  Describe  the  graph  of  g(x,  y,  z)  =  x2  +  y2  by 

computing  some  level  surfaces. 

(b)  Suppose  g  is  a  function  such  that  the  expres- 
sion for  g(x,  y,  z)  involves  only  x  and  y  (i.e., 
g(x,  y,  z)  =  h(x,  yj).  What  can  you  say  about  the 
level  surfaces  of  g? 

(c)  Suppose  g  is  a  function  such  that  the  expression 
for  g(x,  y,  z)  involves  only  x  and  z.  What  can  you 
say  about  the  level  surfaces  of  g? 

(d)  Suppose  g  is  a  function  such  that  the  expression 
for  g(x,  y,  z)  involves  only  x.  What  can  you  say 
about  the  level  surfaces  of  g? 

38.  This  problem  concerns  the  surface  determined  by  the 
graph  of  the  equation  x2  +  xy  —  xz  =  2. 

(a)  Find  a  function  F(x,  y,  z)  of  three  variables  so 
that  this  surface  may  be  considered  to  be  a  level 
set  of  F. 

(b)  Find  a  function  f(x,  y)  of  two  variables  so  that 
this  surface  may  be  considered  to  be  the  graph  of 
z  =  fix,  y). 


42.  x 


y2  z2 


39.  Graph  the  ellipsoid 


x2      y2  2 


i. 


Is  it  possible  to  find  a  function  f(x ,  y )  so  that  this  ellip- 
soid may  be  considered  to  be  the  graph  of  z  =  f(x,  y)? 
Explain. 

Sketch  or  describe  the  surfaces  in  R3  determined  by  the  equa- 
tions in  Exercises  40—46. 


2  2 

43.  x2  +  —  -  —  =  0 
9  16 


40.  z 


y 


41.  z2 


y 


x        yi  z" 

44.  —  +  —  =  1 

4      16  9 

x2      y2  - 

45.  —  +  —  =  z2  -  1 
25  16 

46.  z  =  y2  +  2 

We  can  look  at  examples  of  quadric  surfaces  with  centers 
or  vertices  at  points  other  than  the  origin  by  employing  a 
change  of  coordinates  of  the  form  x  =  x  —  xq,  y  =  y  —  y0, 
and  z  =  Z  —  Zo-  This  coordinate  change  simply  puts  the  point 
(xo,  yo,  Zo)  of  the  xyz-coordinate  system  at  the  origin  of  the 
xyz-coordinate  system  by  a  translation  of  axes.  Then,  for  ex- 
ample, the  surface  having  equation 

(x  -  l)2     (y  +  2)2  , 

can  be  identified  by  setting  x  =  x  —  1,  y  =  y  +  2,  and  z  = 
Z  —  5,  so  that  we  obtain 

-2  -2 

X         V  _2 

which  is  readily  seen  to  be  an  ellipsoid  centered  at  (1,  —2,  5) 
of  the  xyz-coordinate  system.  By  completing  the  square  in  x, 
y,  or  z  as  necessary,  identify  and  sketch  the  quadric  surfaces 
in  Exercises  47—52. 

47.  (*  -  l)2  +  (y  +  l)2  =  (z  +  3)2 

48.  z  =  4*2  +  (y  +  2)2 

49.  Ax2  +  y2  +  z2  +  8*  =  0 

50.  Ax2  +  y2  -  4z2  +  8x  -  4y  +  4  =  0 

51.  *2  +  2y2-6jt-z+10  =  0 

52.  9x2  +  4y2  -  36z2  -  8y  -  144z  =  104 


2.2  Limits 

As  you  may  recall,  limit  processes  are  central  to  the  development  of  calculus.  The 
mathematical  and  philosophical  debate  in  the  1 8th  and  1 9th  centuries  surrounding 
the  meaning  and  soundness  of  techniques  of  taking  limits  was  intense,  questioning 
the  very  foundations  of  calculus.  By  the  middle  of  the  19th  century,  the  infamous 
"e  —  8"  definition  of  limits  had  been  devised  chiefly  by  Karl  Weierstrass  and 
Augustin  Cauchy,  much  to  the  chagrin  of  many  20th  (and  2 1  st)  century  students  of 
calculus.  In  the  ensuing  discussion,  we  study  both  the  intuitive  and  rigorous  mean- 
ings of  the  limit  of  a  function  f :  X  C  R"  — >  Rm  and  how  limits  lead  to  the  notion 
of  a  continuous  function,  our  main  object  of  study  for  the  remainder  of  this  text. 


98       Chapter  2  I  Differentiation  in  Several  Variables 


The  Notion  of  a  Limit   

For  a  scalar- valued  function  of  a  single  variable,  /:XcR->R,  you  have  seen 
the  statement 

lim  f(x)  =  L 

and  perhaps  have  an  intuitive  understanding  of  its  meaning.  In  imprecise  terms, 
the  preceding  equation  (read  "The  limit  of  fix)  as  x  approaches  a  is  L.")  means 
that  you  can  make  the  numerical  value  of  f(x)  arbitrarily  close  to  L  by  keeping 
x  sufficiently  close  (but  not  equal)  to  a.  This  idea  generalizes  immediately  to 
functions  f:Xc  R"  — ►  Rm .  In  particular,  by  writing  the  equation 

lim  f(x)  =  L, 

x^a 

where  f:Xc  R"  Rm ,  we  mean  that  we  can  make  the  vector  f(x)  arbitrarily 
close  to  the  limit  vector  L  by  keeping  the  vector  x  g  X  sufficiently  close  (but  not 
equal)  to  a. 

The  word  "close"  means  that  the  distance  (in  the  sense  of  §  1 .6)  between  f(x) 
and  L  is  small.  Thus,  we  offer  a  first  definition  of  limit  using  the  notation  for 
distance. 


DEFINITION  2.1    (Intuitive  definition  of  limit)  The  equation 

lim  f(x)  =  L, 

x— »a 

where  HcR"^  Rm,  means  that  we  can  make  ||f(x)  —  L||  arbitrarily 
small  (i.e.,  near  zero)  by  keeping  ||x  —  a||  sufficiently  small  (but  nonzero). 


In  the  case  of  a  scalar- valued  function  /:XcR"  ->  R,  the  vector  length 
||f(x)  —  L||  can  be  replaced  by  the  absolute  value  |/(x)  —  L\.  Similarly,  if  /  is  a 
function  of  just  one  variable,  then  ||x  —  a||  can  be  replaced  by  \x  —  a\. 


EXAMPLE  1    Suppose  that  /:  R  ->  R  is  given  by 


1 


Figure  2.30  The  graph  of  /  of 
Example  1. 


/(*)  = 


0 


ifx  <  1 
if  x  >  1 


The  graph  of  /  is  shown  in  Figure  2.30.  What  should  limv^i  f(x)  be?  The  limit 
can't  be  0,  because  no  matter  how  near  we  make  x  to  1  (i.e.,  no  matter  how  small 
we  take  \x  —  1 1),  the  values  of  x  can  be  both  slightly  larger  and  slightly  smaller 
than  1 .  The  values  of  /  corresponding  to  those  values  of  x  larger  than  1  will 
be  2.  Thus,  for  such  values  of  x,  we  cannot  make  \f(x)  —  0|  arbitrarily  small, 
since,  for  x  >  1,  \f(x)  —  0|  =  |2  —  0|  =  2.  Similarly,  the  limit  can't  be  2,  since 
no  matter  how  small  we  take  \x  —  l\,x  can  be  slightly  smaller  than  1.  For  x  <  1, 
f(x)  =  0  and  therefore,  we  cannot  make  \f(x)  —  2|  =  |0  —  2|  =  2  arbitrarily 
small.  Indeed  it  should  now  be  clear  that  the  limit  can't  be  L  for  any  L  e  R. 
Hence,  limx^i  f(x)  does  not  exist  for  this  function.  ♦ 


EXAMPLE  2  Let  f:  R2  R2  be  defined  by  f(x)  =  5x.  (That  is,  f  is  five  times 
the  identity  function.)  Then  it  should  be  obvious  intuitively  that 


lim  f(x) 

x^i+j 


lim  5x : 


5i  +  5j. 


2.2  |  Limits 


Indeed,  if  we  write  x  =  xi  +  yj,  then 
||f(x)-(5i  +  5j)||  =  ||(5xi  +  5yj)-(5i  +  5j)|| 

=  \\5(x-  l)i  +  5(y-  l)j||  =  V25(x-  l)2  +  25(j-  l)2 
=  5j(x  -  l)2  +  (y  -  1)2. 

This  last  quantity  can  be  made  as  small  as  we  wish  by  keeping 

I|X  "  (i  +  j)||  =  y/{x  ~  l)2  +  (y  -  l)2 

sufficiently  small.  ♦ 

EXAMPLE  3  Now  suppose  that  g:R"^R"  is  defined  by  g(x)  =  3x.  We 
claim  that,  for  any  a  e  R", 

lim  3x  =  3a. 

x^a 

In  other  words,  we  claim  that  ||3x  —  3a ||  can  be  made  as  small  as  we  like  by 
keeping  ||x  —  a||  sufficiently  small.  Note  that 

||3x-3a||  =  ||3(x-a)||  =3||x-a||. 

This  means  that  if  we  wish  to  make  ||3x  —  3a||  no  more  than,  say,  0.003,  then 
we  may  do  so  by  making  sure  that  ||x  —  a||  is  no  more  than  0.001.  If,  instead, 
we  want  ||3x  —  3a||  to  be  no  more  than  0.0003,  we  can  achieve  this  by  keeping 
||  x  —  a||  no  more  than  0.0001.  Indeed,  if  we  want  ||3x  —  3a||  to  be  no  more  than 
any  specified  amount  (no  matter  how  small),  then  we  can  achieve  this  by  making 
sure  that  ||x  —  a||  is  no  more  than  one-third  of  that  amount. 

More  generally,  if  h:  R"  —>  R"  is  any  constant  k  times  the  identity  function 
(i.e.,  h(x)  =  kx)  and  a  e  R"  is  any  vector,  then 

lim  h(x)  =  lim  kx  =  ka.  ♦ 

x^a  x— >a 

The  main  difficulty  with  Definition  2.1  lies  in  the  terms  "arbitrarily  small" 
and  "sufficiently  small."  They  are  simply  too  vague.  We  can  add  some  precision 
to  our  intuition  as  follows:  Think  of  applying  the  function  f:  X  c  R"  R"!  as 
performing  some  sort  of  scientific  experiment.  Letting  the  variable  x  take  on 
a  particular  value  in  X  amounts  to  making  certain  measurements  of  the  input 
variables  to  the  experiment,  and  the  resulting  value  f(x)  can  be  considered  to  be  the 
outcome  of  the  experiment.  Experiments  are  designed  to  test  theories,  so  suppose 
that  this  hypothetical  experiment  is  designed  to  test  the  theory  that  as  the  input  is 
closer  and  closer  to  a,  then  the  outcome  gets  closer  and  closer  to  L.  To  verify  this 
theory,  you  should  establish  some  acceptable  (absolute)  experimental  error  for  the 
outcome,  say,  0.05.  That  is,  you  want  ||f(x)  —  L||  <  0.05,  if  ||x  —  a||  is  sufficiently 
small.  Then  just  how  small  does  ||x  —  a  |  need  to  be?  Perhaps  it  turns  out  that  you 
must  have  ||x  —  a||  <  0.02,  and  that  if  you  do  take  ||x  —  a||  <  0.02,  then  indeed 
||f(x)  —  L||  <  0.05.  Does  this  mean  that  your  theory  is  correct?  Not  yet.  Now, 
suppose  that  you  decide  to  be  more  exacting  and  will  only  accept  an  experimental 
error  of  0.005  instead  of  0.05.  In  other  words,  you  desire  ||f(x)  —  L||  <  0.005. 
Perhaps  you  find  that  if  you  take  ||x  —  a||  <  0.001,  then  this  new  goal  can  be 
achieved.  Is  your  theory  correct?  Well,  there's  nothing  sacred  about  the  number 
0.005,  so  perhaps  you  should  insist  that  ||f(x)  —  L||  <  0.001,  or  that  ||f(x)  —  L||  < 
0.00001.  The  point  is  that  if  your  theory  really  is  correct,  then  no  matter  what 
(absolute)  experimental  error  e  you  choose  for  your  outcome,  you  should  be 
able  to  find  a  "tolerance  level"  8  for  your  input  x  so  that  if  ||x  —  a||  <  <5,  then 


Chapter  2  |  Differentiation  in  Several  Variables 


|| f(x)  —  L ||  <  e.  It  is  this  heuristic  approach  that  motivates  the  technical  definition 
of  the  limit. 


DEFINITION  2.2    (RIGOROUS  DEFINITION  OF  limit)  Let  f :  X  C  R"  Rm 

be  a  function.  Then  to  say 

lim  f(x)  =  L 

x— *a 

means  that  given  any  e  >  0,  you  can  find  a  S  >  0  (which  will,  in  general, 
depend  on  e)  such  that  if  x  e  X  and  0  <  ||x  —  a||  <  <5,  then  ||f(x)  —  L||  <  e. 


The  condition  0  <  ||x  —  a||  simply  means  that  we  care  only  about  values 
f(x)  when  x  is  near  a,  but  not  equal  to  a.  Definition  2.2  is  not  easy  to  use  in 
practice  (and  we  will  not  use  it  frequently).  Moreover,  it  is  of  little  value  insofar 
as  actually  evaluating  limits  of  functions  is  concerned.  (The  evaluation  of  the 
limit  of  a  function  of  more  than  one  variable  is,  in  general,  a  difficult  task.) 

EXAMPLE  4  So  that  you  have  some  feeling  for  working  with  Definition  2.2, 
let's  see  rigorously  that 

lim       (3x  -  5y  +  2z)  =  12 

tw)->-  (1,-1,2) 

(as  should  be  "obvious").  This  means  that  given  any  number  e  >  0,  we  can  find 
a  corresponding  S  >  0  such  that 

ifO  <  \\(x,y,z)-(l,  -1,2)||  <  8,    then  \3x  -5y  +  2z-  12|  <  e. 

(Note  the  uses  of  vector  lengths  and  absolute  values.)  We'll  present  a  formal 
proof  in  the  next  paragraph,  but  for  now  we'll  do  the  necessary  background 
calculations  in  order  to  provide  such  a  proof.  First,  we  need  to  rewrite  the  two 
inequalities  in  such  a  way  as  to  make  it  more  plausible  that  the  e-inequality  could 
arise  algebraically  from  the  S -inequality.  From  the  definition  of  vector  length,  the 
(5-inequality  becomes 

0  <  y/(x-  l)2  +  (y  +  l)2  +  (z-2)2  <  8. 
If  this  is  true,  then  we  certainly  have  the  three  inequalities 

V(jc  -  l)2  =  \x  -  1|  <S, 

y(V+l)2  =  |y+l|  <S, 

y/(z  -  2)2  =  \z-2\  <  S. 

Now,  rewrite  the  left  side  of  the  e-inequality  and  use  the  triangle  inequality  (2) 
of  §1.6: 

\3x  -5y  +  2z-  12|  =  |3(*  -  1)  -  5(y  +  1)  +  2(z  -  2)| 

<  |3(x  -  1)|  +  \5(y  +  1)|  +  \2(z  -  2)| 
=  3\x  -  1|  +5|y  +  1|  +2|z-2|. 

Thus,  if 

0<  \\(x,y,z)-  (1,-1,  2)||  <S, 

then 

\x-  1|  <8,     \y+l\  <S,    and    \z-2\  <  8, 

so  that 


\3x-5y  +  2z-  12|  <  3\x  -  1|  +  5\y  +  1|  +  2\z  -  2\ 
<  38  +  58  +  28  =  108. 


2.2  |  Limits  101 


If  we  think  of  S  as  a  positive  quantity  that  we  can  make  as  small  as  desired,  then 
105  can  also  be  made  small.  In  fact,  it  is  10<5  that  plays  the  role  of  e. 

Now  for  a  formal,  "textbook"  proof:  Given  any  e  >  0,  choose  8  >  0  so  that 
S  <  e/10.  Then,  if 

0<  \\(x,y,z)-  (1,-1, 2)||  <S, 

it  follows  that 

|x-l|<<5,     |y+l|<<5,    and  \z-2\<8, 

so  that 

\3x-5y  +  2z-  12|  <  3|x  -  1|  +  5\y  +  1|+  2|z  -  2| 
<  3<5  +  55  +  25 

=  105  <  10—  =  e. 
~  10 

Thus,  lim(x,VjZ)^(1,_ii2)(3x  —  5y  +  2z)  =  12,  as  desired.  ♦ 

Using  the  same  methods  as  in  Example  4,  you  can  show  that 

lim(fliJfi  +  (22*2  +  ■  ■  ■  +  anxn)  =  a\b\  +  02^2  +  ■  ■  ■  +  a„bn 

for  any  cij,  i  =  1,2,  n. 

Some  Topological  Terminology  

Before  discussing  the  geometric  meaning  of  the  limit  of  a  function,  we  need  to  in- 
troduce some  standard  terminology  regarding  sets  of  points  in  R" .  The  underlying 
geometry  of  point  sets  of  a  space  is  known  as  the  topology  of  that  space. 

Recall  from  §2.1  that  the  vector  equation  ||x  —  a||  =  r,  where  x  and  a  are 
in  R3  and  r  >  0,  defines  a  sphere  of  radius  r  centered  at  a.  If  we  modify  this 
equation  so  that  it  becomes  the  inequality 

Hx-a||<r,  (1) 

then  the  points  x  e  R3  that  satisfy  it  fill  out  what  is  called  a  closed  ball  shown  in 
Figure  2.31.  Similarly,  the  strict  inequality 

||x-a||<r  (2) 

describes  points  x  e  R3  that  are  a  distance  of  less  than  r  from  a.  Such  points 
determine  an  open  ball  of  radius  r  centered  at  a,  that  is,  a  solid  ball  without  the 
boundary  sphere. 

There  is  nothing  about  the  inequalities  (1)  and  (2)  that  tie  them  to  R3.  In  fact, 
if  we  take  x  and  a  to  be  points  of  R",  then  (1)  and  (2)  define,  respectively,  closed 
and  open  n -dimensional  balls  of  radius  r  centered  at  a.  While  we  cannot  draw 
sketches  when  n  >  3,  we  can  see  what  (1)  and  (2)  mean  when  n  is  1  or  2.  (See 
Figures  2.32  and  2.33.) 


y  y 


X 


Figure  2.32  The  closed  and  open  balls  (disks)  in  R2  defined  by  [|x  —  a[|  <  r  and 


Chapter  2  |  Differentiation  in  Several  Variables 


X 


Figure  2.34  The  graph  of  X. 


-s^ — 

Figure  2.37  The  set  X  of 

Example  7. 


Figure  2.33  The  closed  and  open  balls  (intervals)  in  R 
defined  by  \x  —  a\  <  r  and  \x  —  a\  <  r. 


DEFINITION  2.3  A  set  X  C  R"  is  said  to  be  open  in  R"  if,  for  each  point 
x  e  X,  there  is  some  open  ball  centered  at  x  that  lies  entirely  within  X.  A 
point  x  e  R"  is  said  to  be  in  the  boundary  of  a  set  X  c  R"  if  every  open  ball 
centered  at  x,  no  matter  how  small,  contains  some  points  that  are  in  X  and 
also  some  points  that  are  not  in  X.  A  set  X  cR"  is  said  to  be  closed  in  R" 
if  it  contains  all  of  its  boundary  points.  Finally,  a  neighborhood  of  a  point 
x  g  X  is  an  open  set  containing  x  and  contained  in  X. 


It  is  an  easy  consequence  of  Definition  2.3  that  a  set  X  is  closed  in  R"  precisely 
if  its  complement  R"  —  X  is  open. 

EXAMPLE  5   The  rectangular  region 

X  =  {(x,  y)  g  R2  |  -1  <  x  <  1,  -1  <  y  <  2} 

is  open  in  R2.  (See  Figure  2.34.)  Each  point  in  X  has  an  open  disk  around  it 
contained  entirely  in  the  rectangle.  The  boundary  of  X  consists  of  the  four  sides 
of  the  rectangle.  (See  Figure  2.35.)  ♦ 


X 


Figure  2.35  Every  open  disk 
about  a  point  on  a  side  of  rectangle 
X  of  Example  5  contains  points  in 
both  X  and  R2  —  X. 


Figure  2.36  The  set  X 

of  Example  6  consists  of 
the  nonnegative 
coordinate  axes. 


EXAMPLE  6  The  set  X  consisting  of  the  nonnegative  coordinate  axes  in  R3  in 
Figure  2.36  is  closed  since  the  boundary  of  X  is  just  X  itself.  ♦ 

EXAMPLE  7  Don't  be  fooled  into  thinking  that  sets  are  always  either  open  or 
closed.  (That  is,  a  set  is  not  a  door.)  The  set 

X  =  {(x,  y)  g  R2  |  0  <  x  <  1,  0  <  y  <  1} 

shown  in  Figure  2.37  is  neither  open  nor  closed.  It's  not  open  since,  for  example, 
the  point  Q,  0)  that  lies  along  the  bottom  edge  of  X  has  no  open  disk  around  it 
that  lies  completely  in  X.  Furthermore,  X  is  not  closed  since  the  boundary  of  X 
includes  points  of  the  form  (x,  1)  for  0  <  x  <  1  (why?),  which  are  not  part  of  X. 


2.2  |  Limits  103 


The  Geometric  Interpretation  of  a  Limit   

Suppose  that  He  R"  — »  Rm.  Then  the  geometric  meaning  of  the  statement 

lim  f(x)  =  L 

\  Ml 

is  as  follows:  Given  any  e  >  0,  you  can  find  a  corresponding  S  >  0  such  that  if 
points  x  e  X  are  inside  an  open  ball  of  radius  8  centered  at  a,  then  the  correspond- 
ing points  f(x)  will  remain  inside  an  open  ball  of  radius  e  centered  at  L.  (See 
Figure  2.38.) 

y 


Figure  2.38  Definition  of  a  limit:  Given  an  open  ball  5e  centered  at  L  (right),  you 
can  always  find  a  corresponding  ball  Bg  centered  at  a  (left),  so  that  points  in  Bs  n  X 
are  mapped  by  f  to  points  in  Be . 

We  remark  that  for  this  definition  to  make  sense,  the  point  a  must  be  such 
that  every  neighborhood  of  it  in  R"  contains  points  xeX  distinct  from  a.  Such 
a  point  a  is  called  an  accumulation  point  of  X.  (Technically,  this  assumption 
should  also  be  made  in  Definition  2.2.)  A  point  a  e  X  is  called  an  isolated  point 
of  X  if  it  is  not  an  accumulation  point,  that  is,  if  there  is  some  neighborhood  of  a 
in  R"  containing  no  points  of  X  other  than  a. 

From  these  considerations,  we  see  that  the  statement  limx^a  f(x)  =  L  really 
does  mean  that  as  x  moves  toward  a,  f(x)  moves  toward  L.  The  significance  of 
the  "open  ball"  geometry  is  that  entirely  arbitrary  motion  is  allowed. 

EXAMPLE  8   Let  /:  R2  -  {(0,  0)}  ->  R  be  defined  by 


/(*.  y)  = 


y 


x2  +  y2 

Let's  see  what  happens  to  /  as  x  =  (x,  y)  approaches  0  =  (0,  0).  (Note  that  /  is 
undefined  at  the  origin,  although  this  is  of  no  consequence  insofar  as  evaluating 
limits  is  concerned.)  Along  the  x-axis  (i.e.,  the  line  y  =  0),  we  calculate  the  value 
of  /  to  be 

,•2 


f(x,  0)  = 


0 


0 


1. 


Thus,  as  x  approaches  0  along  the  line  y  =  0,  the  values  of  /  remain  constant, 
and  so 

lim  /(x)=l. 

x— >  0  along  y=0 

Along  the  j-axis,  however,  the  value  of  /  is 


m  y) 


0-y2 
0  +  y2 


1  04       Chapter  2  |  Differentiation  in  Several  Variables 


Hence, 


lim       /(x)  =  -l. 

>0  along  x=0 


Indeed,  the  value  of  /  is  constant  along  each  line  through  the  origin.  Along  the 
line  y  =  mx,  m  constant,  we  have 


f(x,  mx) 

Therefore, 


x2  —  m2x2      x2(l  —  m2)      1  —  m2 


x2+m2x2      x2(\  +  m2)      1  +  m2 


1  -  m2 

lim  f(x) 


x^O  along  y=mx  1  -|-  m 

As  a  result,  the  limit  of  /  as  x  approaches  0  does  not  exist,  since  /  has  different 
"limiting  values"  depending  on  which  direction  we  approach  the  origin.  (See 
Figure  2.39.)  That  is,  no  matter  how  close  we  come  to  the  origin,  we  can  find 
points  x  such  that  f(x)  is  not  near  any  number  LeR,  (In  other  words,  every 
open  disk  centered  at  (0,  0),  no  matter  how  small,  is  mapped  onto  the  interval 
[  —  1 ,  1].)  If  we  graph  the  surface  having  equation 

2  2 

x  —  y 
x2  +  y2 

(Figure  2.40),  we  can  see  quite  clearly  that  there  is  no  limiting  value  as  x  ap- 
proaches the  origin.  ♦ 


y 

R2 


Figure  2.39  The  function  f{x,  y)  =  (x2  -  y2)/(x2  +  y2)  of  Figure  2.40  The  graph  of  f(x,  y)  = 

Example  8  has  value  1  along  the  jc-axis  and  value  - 1  along  the  (*  -  y  )/(*  +  y1)  of  Example  8. 

y-axis  (except  at  the  origin). 


Warning  Example  8  might  lead  you  to  think  you  can  establish  that 
lim^a  f(x)  =  L  by  showing  that  the  values  of  f  as  x  approaches  a  along  straight- 
line  paths  all  tend  toward  the  same  value  L.  Although  this  is  certainly  good 
evidence  that  the  limit  should  be  L,  it  is  by  no  means  conclusive.  See  Exercise  23 
for  an  example  that  shows  what  can  happen. 

EXAMPLE  9  Another  way  we  might  work  with  the  function  f(x,  y)  =  (x2  — 
y2)/(x2  +  y2)  of  Example  8  is  to  rewrite  it  in  terms  of  polar  coordinates.  Thus, 
let  x  =  r  cos  9,  y  =  r  sin  9.  Using  the  Pythagorean  identity  and  the  double  angle 


2.2  |  Limits  105 

formula  for  cosine,  we  obtain,  for  r  ^  0,  that 

x2  -  y2      r2  cos2  9  -  r2  sin2  9      r2(cos2  9  -  sin2  9)      cos  29 

 =  ~  =  —  cos  2$ 

x2  +  y2      r2  cos2  (9  +  r2  sin2  (9      r2(cos2  6»  +  sin2  6»)  1 

That  is,  for  r  /  0, 

f(x,y)  =  f(r  cos0,  r  sin#)  =  cos 20. 

Moreover,  to  evaluate  the  limit  of  /  as  (x,  y)  approaches  (0,  0),  we  only  must 
have  r  approach  0;  there  need  be  no  restriction  on  9.  Therefore,  we  have 

lim     f(x,  y)  =  limcos2#  =  cos2#. 

This  result  clearly  depends  on  9.  For  example,  if  9  =  0  (which  defines  the  x-axis), 
then 

lim       cos  29  =  1, 

r^O  along  6  =  0 

while  if  9  =  tt/4  (which  defines  the  line  y  =  x),  then 

lim        cos  29  =  0. 

r^0  along  6  =  jr/4 

Thus,  as  in  Example  8,  we  see  that  lim(A  V)^(o,o)  f(x,  y)  fails  to  exist.  ♦ 

EXAMPLE  10  We  use  polar  coordinates  to  investigate  lim(A>Y)^(o,o)  f(x,  y), 
where  f(x ,  y)  =  (x3  +  x5)/(x2  +  y2). 

We  first  rewrite  the  expression  (x3  +  x5)/(x2  +  y2)  using  polar  coordinates: 


x3  +  x5      r3  cos3  0  +  r5  cos5 1 


x2  +  y2      r2  cos2  6>  +  r2  sin 
Now  —  1  <  cos#  <  1,  which  implies  that 


r(cos3  9  +  r2  cos5  9). 


Hence, 


■  1  -  r2  <  cos3  9  +  r2  cos5  9  <  1  +  r2. 


-r(\  +  r2)<  f(x,y)<r(l+r2). 


As  r  ->  0,  both  the  expressions  — r(l  +  r2)  and  r(l  +  r2)  approach  zero.  Hence, 
we  conclude  that  lim(x,v)^(o,o)  /(x,  y)  =  0,  since  /  is  squeezed  between  two 
expressions  with  the  same  limit.  ♦ 

Properties  of  Limits   


One  of  the  biggest  drawbacks  to  Definition  2.2  is  that  it  is  not  at  all  useful  for 
determining  the  value  of  a  limit.  You  must  already  have  a  "candidate  limit"  in 
mind  and  must  also  be  prepared  to  confront  some  delicate  work  with  inequalities 
to  use  Definition  2.2.  The  results  that  follow  (which  are  proved  in  the  addendum 


Chapter  2  |  Differentiation  in  Several  Variables 


to  this  section),  plus  a  little  faith,  can  be  quite  helpful  for  establishing  limits,  as 
the  subsequent  examples  demonstrate. 


THEOREM  2.4  (Uniqueness  OF  limits)  If  a  limit  exists,  it  is  unique.  That  is, 
let  f:  X  C  R"      Rm.  If  limx^a  f(x)  =  L  and  limx^a  f(x)  =  M,  then  L  =  M. 


THEOREM  2.5  (Algebraic  properties)  Let  F,G:Xc  R"  ->  R"  be  vector- 
valued  functions,  f,g:X^  R"  ->  R  be  scalar-valued  functions,  and  let  k  €  R 
be  a  scalar. 

1.  If  limx^a  F(x)  =  L  and  limx^a  G(x)  =  M, 
then  limx^a(F  +  G)(x)  =  L  +  M. 

2.  If  linwa  F(x)  =  L,  then  limx^a  k¥(x)  =  kh. 

3.  If  limx^a  f(x)  =  L  and  limx^a  g(x)  =  M,  then  limx^a(/g)(x)  =  LM. 

4.  If  limx^a  /(x)  =  L,  g(x)  /  0  for  x  e  X,  and  limx^a  g(x)  =  M/  0,  then 
linwa(//g)(x)  =  L/M. 


There  is  nothing  surprising  about  these  theorems — they  are  exactly  the  same 
as  the  corresponding  results  for  scalar- valued  functions  of  a  single  variable.  More- 
over, Theorem  2.5  renders  the  evaluation  of  many  limits  relatively  straightforward. 

EXAMPLE  1 1  Either  from  rigorous  considerations  or  blind  faith,  you  should 
find  it  plausible  that 

lim     x  =  a    and        lim     v  =  b. 

(x,y)^(a,b)  (x,y)^(a,b) 

From  these  facts,  it  follows  from  Theorem  2.5  parts  1,  2,  and  3  that 

lim    (x2  +  2xy  -  y3)  =  a2  +  lab  -  b3 , 

(x,y)-*(a,b) 

because,  by  part  1  of  Theorem  2.5, 

lim    (x2  +  2xy  -  y3)  =  limx2  +  lim2jcy  +  lim(-y3) 

(x,y)-^(a,b) 

and,  by  parts  2  and  3, 

lim    (x2  +  Ixy  -  y3)  =  (limx)2  +  2(limx)(limy)  -  (limy)3 

(x,y)-*(a,b) 

so  that,  from  the  facts  just  cited, 

lim     (x2  +  2xy  -  y3)  =  a2  +  lab  -  b3 .  ♦ 

(x,yy+(a,b) 

EXAMPLE  12  More  generally,  a  polynomial  in  two  variables  x  and  y  is  any 
expression  of  the  form 

d  d 

p(x,y)  =  J2J2ckixkyl, 

k=0  1=0 

where  d  is  some  nonnegative  integer  and  c«  6  R  for  k,  I  =  0, . . . ,  d.  That  is, 
p(x,  y)  is  an  expression  consisting  of  a  (finite)  sum  of  terms  that  are  real  number 
coefficients  times  powers  of  x  and  y.  For  instance,  the  expression  x2  +  2xy  —  y3 
in  Example  11  is  a  polynomial.  For  any  (a,  b)  e  R2,  we  have,  by  part  1  of 


2.2  |  Limits  107 


Theorem  2.5, 

d  d 

lim     p(x,  v)  =  y^y^     lim  (ctixkyl) 
so  that,  from  part  2, 

d  d 

lim     p(x,  y)  =  7^  7^ cm     lim  xkyl 

(x,y)->(a,b)  f^0  (x,y)^(a,b) 

and,  from  part  3, 


d  d 

lim     p(x,  j)  =  VV  cH(limx*)(limyz) 
=  ±±cklakbK 

k=0  1=0 

Similarly,  a  polynomial  in  n  variables  x\,  x%, . . . ,  xn  is  an  expression  of  the  form 


p(x\,x2, . . .  ,xn)  =    ^    c*:i..4nxf1x22  •  •  •**" 


fei,...,ft„=0 

where  is  some  nonnegative  integer  and  c*,...^  6  R  for  fci, . . . ,  fc„  =  0, . . . ,  d. 
For  example,  a  polynomial  in  four  variables  might  look  like  this: 

p{X\,  .  .  .  ,  Xa)  =  1x\x-i  +  —  7X3X4. 

Theorem  2.5  implies  readily  that 

lim  -k,  X\  x2  ■■■  xn"  =       Ckv-Ka\  a2  ' ' '  a„  ■ 

EXAMPLE  13   We  evaluate      lim  *  +Xy  +  ]  . 

(x,yy+(-l,0)  x2y  -  5xy  +  y2  +  1 

Using  Example  12,  we  see  that 

lim      x2  +  xv  +  3  =  4, 

(x,yH-(-l,0) 

and 

lim     x2y  -  5xy  +  y2  +  1  =  1(/  0). 

(A-,y)^<-1.0)  ' 

Thus,  from  part  4  of  Theorem  2.5,  we  conclude  that 

lim  *2  +  '?  +  3       =1  =  4. 

(x,y)^(-i,0)  x2y  -  5xy  +  y2  +  1  1 

EXAMPLE  14  Of  course,  not  all  limits  of  quotient  expressions  are  as  simple 
to  evaluate  as  that  of  Example  13.  For  instance,  we  cannot  use  Theorem  2.5  to 
evaluate 

2  4 
x  —  y 

lim     —  j  (3) 

since  ]jm(Xtyy+(o,o)(x2  +  y4)  =  0.  Indeed,  since  lim^^-^o.o^x2  —  y4)  =  0  as 
well,  the  expression  (x2  —  y4)/(x2  +  y4)  becomes  indeterminate  as  (x,  y)  -> 
(0,  0).  To  see  what  happens  to  the  expression,  we  note  that 

x2      4  x2 
lim       —          =  lim  —  =  1, 

.1-^0  along  y=0  X2  +  y4       x-*0  X2 


Chapter  2  |  Differentiation  in  Several  Variables 


while 


lim 

>0  along 


=o  x2  +  y4 


=  lim 


v4 


Thus,  the  limit  in  (3)  does  not  exist.  (Compare  this  with  Example  8.)  ♦ 

The  following  result  shows  that  evaluating  the  limit  of  a  function 
f :  X  c  R"  —>  R"  is  equivalent  to  evaluating  the  limits  of  its  (scalar- valued)  com- 
ponent functions.  First  recall  from  §2.1  that  f(x)  may  be  rewritten  as  (/i(x), 
/2(x),  . . . ,  /m(x)). 

THEOREM  2.6  Suppose  He  R"  R"1  is  a  vector-valued  function.  Then 
limx^a  f(x)  =  L,  where  L  =  (Li,  . . . ,  Lm),  if  and  only  if  limx^a  ft(x)  =  Li  for 
i  =  1,  ....  m. 


EXAMPLE  15  Consider  the  linear  mapping  f:  R"  -■>•  R"  defined  by  f(x)  =  Ax, 
where  A  =  (a,;)  is  an  m  x  n  matrix  of  real  numbers.  (See  Example  5  of  §1.6.) 
Theorem  2.6  shows  us  that 


for  any  h  =  (b\ 


lim  f(x)  =  Ab 

x— >b 

,  bn)  in  R" .  If  we  write  out  the  matrix  multiplication,  we  have 


f(x)  =  Ax  = 


flu 

«21 
_  Ami 


flln 
fl2;i 


-Vl 

A'2 


mn  _1  l_  -^n  _ 


aUX\  +  (312^2  + 
Cl2\X\  +  a22*2  + 


~f"  Cl\n%n 


flml-^-1  ~T~  @m2X2   1***1  0.mnXn 

Therefore,  the  tth  component  function  of  f  is 

=  anxi  +  a,-2x2  H  h  fli«x„. 

From  Example  4,  we  have  that 

lim  yi(x)  =  fln^i  +  fl/2^2  H  h  a,>A 

x^b 

for  each  /.  Hence,  Theorem  2.6  tells  us  that  the  limits  of  the  component  functions 
fit  together  to  form  a  limit  vector.  We  can,  therefore,  conclude  that 


limf(x)  =  (lim/1(x), . 

x^b  x^b 

=  (aubi  H  

anb]  + 
a2\b\  + 


,  lim  /m(x)) 

\  >  b 

■  ci\nbn, . . . ,  am\b\  + 

+  a\nbn 
+  a2nbn 


®mnb\\) 


am\b\  +  •  •  ■  +  a„,„b„ 


once  we  take  advantage  of  matrix  notation. 


Ab, 


2.2  |  Limits  109 


Figure  2.41  The  graph  of  a 
continuous  function. 


Figure  2.44  The  graph  of  a 
continuous  function  f(x,  y). 


Continuous  Functions   

For  scalar-valued  functions  of  a  single  variable,  one  often  adopts  the  following 
attitude  toward  the  notion  of  continuity:  A  function  /:  X  C  R  is  continuous 
if  its  graph  can  be  drawn  without  taking  the  pen  off  the  paper.  By  this  criterion, 
Figure  2.41  describes  a  continuous  function  y  =  f(x),  while  Figure  2.42  does  not. 


Figure  2.42  The  graph  of  a 
function  that  is  not  continuous. 


Figure  2.43  The  graph  of  / 
where  f(x,  y)  =  0  if  both  x  >  0 
and  y  >  0,  and  where  f(x,  y)  =  1 
otherwise. 


We  can  try  to  extend  this  idea  to  scalar-valued  functions  of  two  variables:  A 
function  /:  X  C  R2  — R  is  continuous  if  its  graph  (in  R3)  has  no  breaks  in  it. 
Then  the  function  shown  in  Figure  2.43  fails  to  be  continuous,  but  Figure  2.44 
depicts  a  continuous  function.  Although  this  graphical  approach  to  continuity  is 
pleasantly  geometric  and  intuitive,  it  does  have  real  and  fatal  flaws.  For  one  thing, 
we  can't  visualize  graphs  of  functions  of  more  than  two  variables,  so  how  will  we 
be  able  to  tell  in  general  if  a  function  /:XcR"^  Rm  is  continuous?  Moreover, 
it  is  not  always  so  easy  to  produce  a  graph  of  a  function  of  two  variables  that  is 
sufficient  to  make  a  visual  determination  of  continuity.  This  said,  we  now  give  a 
rigorous  definition  of  continuity  of  functions  of  several  variables. 


DEFINITION  2.7  Let  f:  X  C  R"  ->  R"!  and  let  a  e  X.  Then  f  is  said  to  be 
continuous  at  a  if  either  a  is  an  isolated  point  of  X  or  if 

limf(x)  =  f(a). 

x->a 

If  f  is  continuous  at  all  points  of  its  domain  X,  then  we  simply  say  that  f  is 
continuous. 


EXAMPLE  16   Consider  the  function  /:  R2 

x2  +  xy  —  2y2 

f(x,y)=  ' 


x2  +  y2 


0 


->  R  defined  by 
if  (x,  y)  #  (0,  0) 
if  (x,  y)  =  (0,  0) 


Therefore,  f(0,  0)  =  O^utlim^^^o)  f(x,  y)  does  not  exist.  (To  see  this,  check 
what  happens  as  (x,  y)  approaches  (0,0)  first  along  y  =  0  and  then  along  x  =  0.) 
Hence,  /  is  not  continuous  at  (0,0).  ♦ 

It  is  worth  noting  that  Definition  2.7  is  nothing  more  than  the  "vectorized" 
version  of  the  usual  definition  of  continuity  of  a  (scalar-valued)  function  of  one 
variable.  This  definition  thus  provides  another  example  of  the  power  of  our  vector 
notation:  Continuity  looks  the  same  no  matter  what  the  context. 


Chapter  2  |  Differentiation  in  Several  Variables 


One  way  of  thinking  about  continuous  functions  is  that  they  are  the  ones  whose 
limits  are  easy  to  evaluate:  When  f  is  continuous,  the  limit  of  f  as  x  approaches 
a  is  just  the  value  of  f  at  a.  It's  all  too  tempting  to  get  into  the  habit  of  behaving 
as  if  all  functions  are  continuous,  especially  since  the  functions  that  will  be  of 
primary  interest  to  us  will  be  continuous.  Try  to  avoid  such  an  impulse. 

EXAMPLE  1 7  Polynomial  functions  in  n  variables  are  continuous.  Example  12 
gives  a  sketch  of  the  fact  that 

lim  y*  ckv..kix\1  ■  ■  ■  xl"  =  y~]  ch...kaa^  ■  ■  ■  a*", 

where  x  =  (xi, . . . ,  xn)  and  a  =  (#i, . . . ,  an)  are  in  R".  If  /:  R"  —>  R  is  defined 
by 

/(x)  =  X>*,...fc*?'---#. 
then  the  preceding  limit  statement  says  precisely  that  /  is  continuous  at  a.  ♦ 

EXAMPLE  18  Linear  mappings  are  continuous.  If  f:  R"  — >  R'"  is  denned  by 
f(x)  =  Ax,  where  A  is  an  m  x  n  matrix,  then  Example  1 5  establishes  that 

lim  f(x)  =  Ab  =  f(b) 

for  all  b  e  R".  Thus,  f  is  continuous.  ♦ 

The  geometric  interpretation  of  the  e  —  S  definition  of  a  limit  gives  rise  to  a 
similar  interpretation  of  continuity  at  a  point:  f:  X  C  R"  —>  R"!  is  continuous  at 
a  point  a  £  X  if,  for  every  open  ball  B€  in  R"  of  radius  e  centered  at  f(a),  there 
is  a  corresponding  open  ball  Bs  in  R"  of  radius  8  centered  at  a  such  that  points 
x  e  X  inside  B&  are  mapped  by  f  to  points  inside  B(.  (See  Figure  2.45.)  Roughly 
speaking,  continuity  of  f  means  that  "close"  points  in  X  c  R"  are  mapped  to 
"close"  points  inRffl. 


In  practice,  we  usually  establish  continuity  of  a  function  through  the  use 
of  Theorems  2.5  and  2.6.  These  theorems,  when  interpreted  in  the  context  of 


2.2  |  Limits        1 1  1 


continuity,  tell  us  the  following: 


•  The  sum  F  +  G  of  two  functions  F,  G:  X  C  R"  — >  Rm  that  are  continuous 
at  a  e  X  is  continuous  at  a. 

•  For  all  &  €  R,  the  scalar  multiple  kF  of  a  function  F:XcR"->Rffl  that 
is  continuous  at  a  £  X  is  continuous  at  a. 

•  The  product  fg  and  the  quotient  f/g(gj^  0)  of  two  scalar- valued  functions 
f,g:XC.  R"      R  that  are  continuous  ata  e  I  are  continuous  at  a. 

•  F:XcR"->  R'"  is  continuous  at  a  e  X  if  and  only  if  its  component 
functions  F* :  X  C  R"  — »•  R,  j  =  1 , . . . ,  m  are  all  continuous  at  a. 


EXAMPLE  1 9   The  function  f:  R2  ->  R3  denned  by 

f(x,  y)  =  (x  +  y,  x2y,  y  sin(xy)) 

is  continuous.  In  view  of  the  remarks  above,  we  can  see  this  by  checking  that  the 
three  component  functions 

fi(x,y)  =  x  +  y,     f2(x,y)  =  x2y,     and    f3(x,  y)  =  y  sin(xy) 

are  each  continuous  (as  scalar-valued  functions).  Now  f\  and  fi  are  continuous, 
since  they  are  polynomials  in  the  two  variables  x  and  y.  (See  Example  17.)  The 
function     is  the  product  of  two  further  functions;  that  is, 

Mx,  y)  =  g(x,  y)h(x,  y), 

where  g{x,  y)  =  y  and  h(x,  y)  =  sin(xy).  The  function  g  is  clearly  continuous. 
(It's  a  polynomial  in  two  variables — one  variable  doesn't  appear  explicitly!)  The 
function  h  is  a  composite  of  the  sine  function  (which  is  continuous  as  a  function 
of  one  variable)  and  the  continuous  function  p(x,  y)  =  xy.  From  these  remarks, 
it's  not  difficult  to  see  that 

lim     h(x,  y)  =     lim     sin(p(x,  y)) 

(x,y)-+(a,b)  (x,y)^(a,b) 

=  sin  (     lim     p(x,  y)  I  , 

since  the  sine  function  is  continuous.  Thus, 

lim     h(x,  y)  =  sin p(a,  b)  =  h(a,  b), 

because  p  is  continuous.  Thus,  h,  hence  fa,  and,  consequently,  fare  all  continuous 
on  all  of  R2.  ♦ 

The  discussion  in  Example  19  leads  us  to  the  following  general  result,  whose 
proof  we  omit: 


THEOREM  2.8  Iff:  XcR"^  R"<  andg:  Y  C  R"  RP  are  continuous  func- 
tions such  that  range  fey,  then  the  composite  function  g  o  f:  X  C  R"  — >  Rp  is 
defined  and  is  also  continuous. 


Chapter  2  |  Differentiation  in  Several  Variables 


Addendum:  Proofs  of  Theorems  2.4,  2.5,  2.6,  and  2.8   

For  the  interested  reader,  we  establish  the  various  results  regarding  limits  of 
functions  that  we  used  earlier  in  this  section. 

Proof  of  Theorem  2.4  The  statement  limx^a  f(x)  =  L  means  that,  given  any  e  > 
0,  we  can  find  some  Si  >  0  such  that  if  x  6  X  and  0  <  ||x  —  a||  <  Si,  then  ||f(x)  — 
L||  <  e/2.  (The  reason  for  writing  e/2  rather  than  e  will  become  clear  in  a 
moment.)  Similarly,  limx^a  f(x)  =  M  means  that,  given  any  e  >  0,  we  can  find 
some  S2  >  0  such  that  if  x  e  X  and  0  <  ||x  -  a||  <  S2,  then  ||f(x)  -  M||  <  e/2. 

Now  let  S  =  min^!,  £2);  that  is,  we  set  S  to  be  the  smaller  of  Si  and  S2.  If 
x  e  X  and  0  <  ||x  -  a||  <  8,  then  both  ||f(x)  -  L||  and  ||f(x)  -  M||  are  less  than 
e /2  so  that,  using  the  triangle  inequality,  we  have 

||L-M||  =  ||(L-f(x))  +  (f(x)-M)|| 

<  ||L-f(x)||  +  ||f(x)-M||  <|  +  |  =  6. 

This  shows  that  the  quantity  ||L  —  M||  can  be  made  arbitrarily  small;  thus,  it 
follows  that  L  -  M  =  0.  Hence,  L  =  M.  ■ 

Proof  of  Theorem  2.5  To  establish  part  1,  note  that  if  limx^aF(x)  =  L,  then 
given  any  e  >  0,  we  can  find  a  Si  >  0  such  that  if  x  e  X  and  0  <  ||x  —  a||  < 
Si,  then  || F(x)  —  L||  <  e/2.  Similarly,  if  limx^a  G(x)  =  M,  then  we  can  find  a 
S2  >  0  such  that  if  x  6  X  and  0  <  ||x  -  a||  <  S2,  then  ||G(x)  -  M||  <  e/2.  Now 
let  i5  =  min((5i,  ^2).  Then  if  x  e  X  and  0  <  ||x  —  a||  <  <5,  the  triangle  inequality 
implies  that 

||(F(x)  +  G(x))  -  (L  +  M)||  <  ||F(x)  -  L||  +  ||G(x)  -  M\\  <"-+"-=  e. 

Hence,  limx^a  (F(x)  +  G(x))  =  L  +  M. 

To  prove  part  2,  suppose  that  e  >  0  is  given.  If  limx^a  F(x)  =  L,  then  we  can 
find  a  <5  >  0  such  that  if  x  £  X  and  0  <  ||x  -  a||  <  S,  then  ||F(x)  -  L||  <  e/\k\. 
Therefore, 

\\k¥(x)  -  *L||  =  |*|  ||F(x)  -  L||  <  |^|  -1  =  e, 

\k\ 

which  means  that  limx^a  k¥(x)  =  kL,.  (Note:  If  k  =  0,  then  part  2  holds  trivially.) 

To  establish  the  rule  for  the  limit  of  a  product  of  scalar- valued  functions  (part 
3),  we  will  use  the  following  algebraic  identity: 

/(x)g(x)  -LM  =  (/(x)  -  L)(g(x)  —  M)  +  L(g(x)  —  M)  +  M(f(x)  -  L). 

(4) 

If  limx^a  f(x)  =  L,  then,  given  any  e  >  0,  we  can  find  Si  >  0  such  that  if  x  e  X 
and  0  <  ||x  —  a||  <  Si,  then 

|/(x)-L|  <  Vi\ 

Similarly,  if  limx^a  g(x)  =  M,  we  can  find  S2  >  0  such  that  if  x  e  X  and  0  < 
||x  —  a||  <  S2,  then 

\g(x)-M\  <  Je. 
Let  8  =  min((5i ,  ^2).  If  x  e  X  and  0  <  ||x  —  a||  <  8,  then 


|(/(x)  -  L)(g(x)  -M)\<sTe-^e  =  e. 


2.2  |  Limits        1 1  3 


This  means  that  limx^a(/(x)  —  L)(g(x)  —  M)  =  0.  Therefore,  using  (4)  and  parts 
1  and  2,  we  see  that 

lim(/(x)g(x)  -  LM)  =  lim(/(x)  -  L)(g(x)  -M)  +  L  lim(g(x)  -  M) 

x^-  a  x->  a  x-^-  a 

+  M  lim(/(x)  -  L) 

x— >a 

=  0  +  0  +  0  =  0. 

Since  limx^a  f(x)g(x)  =  limx^a((/(x)^(x)  —  LM)  +  LM),  the  desired  result 
follows  from  part  1 . 

The  crux  of  the  proof  of  part  4  is  to  show  that 


lim 


1 


1 

~M 


x^a  g(x) 

Once  we  show  this,  the  desired  result  follows  directly  from  part  3 : 


/(x) 


lim 

x^a  g(X) 


Note  that 


=  lim  /(X) 


1 

M 


1 


S(x) 


S(x) 

_  \M  -  g(x)\ 
\Mg(x)\ 


1  _  L 

~      ~M  ~  ~M 


and,  by  the  triangle  inequality,  that 

\M\  =  \M-  g(x)  +  g(x)|  <\M-  g(x)\  +  \g(x) 


(5) 


If  limx^.a  g(x)  =  M,  then,  given  any  e  >  0,  we  can  find  8\  such  that  if  x  e  X  and 
0  <  ||x-  a||  <  <5!,then 

M2 

\g(x)-M\  <—e. 

We  can  also  fmd  82  such  that  if  x  e  X  and  0  <  ||x  —  a||  <  82,  then  |g(x)  —  M\  < 
\M\/2  and,  hence,  using  (5),  that 


|M| 

\M\  <        +  1*001  \g(x)\  > 


\M\ 


1 


Now  let  8  =  min(<5i ,  ^2).  If  x  e  X  and  0 


1 


g(x) 


1 

M 


|M-g(x)| 
|Mg(x)| 
1     2  M2 


2  l#(x)| 
x  —  a||  <  8,  then 

1  |M-g(x)| 
"  IM| 


< 


|M| 


lg(x)l 


<  |M|  |M|   2  e  €' 


Proof  of  Theorem  2.6  Note  first  that,  for  i  =  1, . . . ,  m, 

\fi(x)  ~L,\<  V(/1(x)-L1)2  +  ---  +  (/m(x)-Lm)2  =  ||f(x)  -  L|| 


(6) 


If  limx^a  f(x)  =  L,  then  given  any  e  >  0,  we  can  find  a  5  >  0  such  that  if  x  e 
X  and  0  <  ||x  -  a||  <  8,  then  ||f(x)  -  L||  <  e.  Hence,  (6)  implies  that  |/;(x)  - 
Lf\  <  €  for  i  =  1, . . . ,  m,  which  means  that  limx^a  /,(x)  =  L;. 

Conversely,  suppose  that  limx^a  /i(x)  =  L,-  for  i  =  1, . . . ,  m.  This  means 
that,  given  any  e  >  0,  we  can  find,  for  each  i ,  a  8t  >  0  such  that  if  x  e  X  and 


14       Chapter  2  |  Differentiation  in  Several  Variables 


0  <  ||x  —  a||  <  8i,  then  |/}(x)  —  L,|  <  e/^Jm.  Set  8  =  min(<5i, . . . ,  8m).  Then  if 
x  e  X  and  0  <  ||x  —  a||  <  8,  we  see  that  (6)  implies 

||f(x)-L||  <  J-  +  =    m-  =  €. 

V  m  m      V  m 

Thus,  limx^a  f(x)  =  L.  ■ 

Proof  of  Theorem  2.8  We  must  show  that  the  composite  function  g  o  f  is  con- 
tinuous at  every  point  a  e  X.  If  a  is  an  isolated  point  of  X,  there  is  nothing  to 
show.  Otherwise,  we  must  show  that  limx^a(g  o  f)(x)  =  (go  f)(a). 

Given  any  e  >  0,  continuity  of  g  at  f(a)  implies  that  we  can  find  some  y  >  0 
such  that  if  y  e  range  f  and  0  <  ||y  —  f(a)||  <  y  then 

||g(y)-g(f(a))||  <6. 

Since  f  is  continuous  at  a,  we  can  find  some  <5  >  0  such  that  if  x  e  X  and  0  < 
|| x  —  a||  <  8,  then 

||f(x)-f(a)||  <  y. 
Therefore,  if  x  e  X  and  0  <  ||x  —  a||  <  8,  then 

||g(f(x))-g(f(a))||  <e.  m 

2.2  Exercises 


In  Exercises  1—6,  determine  whether  the  given  set  is  open  or 
closed  (or  neither). 

1.  {(x,y)eR2  |  1  <  x2  +  y2  <  4} 

2.  {(x,y)eR2  |  1  <x2+y2  <  4} 

3.  {(x,y)eR2  |  1  <  x2  +  y2  <  4} 

4.  {(x,  y,  z)  e  R3  |  1  <  x2  +  y2  +  z2  <  4} 

5.  {(x,  y)  e  R2  |  -1  <  x  <  1}  U  {(x,  y)  e  R2  | 

6.  {(x,y,z)eR3  |  1  <  x2  +  y2  <  4} 


2} 


Evaluate  the  limits  in  Exercises  7—21,  or  explain  why  the  limit 
fails  to  exist. 

x2  +  2xy  +  yz  +  z3  +  2 


7.  lim 

(*,.y,z)-H0,0,0) 

8.  lim 


l.v| 


(x,y)->(0,0)  jxl  _|_  yl 

9.      lim     —  - 

(x.y)^r(P.O)  x2  +  y2 

10.  lim 


(jt.yH-ro.O)  x  +  y  +  2 
2x2  +  y2 


1 1 .  lim 


(jt,y)-+(0.0)  X2  +  y2 

2x2  +  y2 


12.  lim 


(x,y)^(-l,2)   X2  +  y2 


x2  +  2xy  +  y2 
u-O  -to  ui        x  +  y 

xy 


13.  lim 

14.  lim 


(*,;>>)-K0,0)  x2  +  y2 

x4-y4 

15.  lim     —  ■— 

(jr,y)-<-(0,0)  x1  +  y1 
X2 

16.  lim     —  - 

(x,y)^(0,0)  X2  +  y2 


17. 


lim 


x  —  xy 


(jc,y)->(o.o)>*54>  -Jx  - 

x2-y2-4x+4 
1 8.      lim     —  —  

(x,y)^(2,0)  x2  +y2  -4x  +4 


19. 

20. 
21. 


lim        exz  cos  y  —  x 

(x,y,z)^(0,^fit.l) 


2x2  +  3y2  +  z2 

(jc,y,t)->-(0,0,a)    X2  +  y2  +  Z2 
xy  —  xz  +  yz 


lim 


lim 


(jc,y,t)-K0,0,a)  x2  +  y2  +  z2 
sin© 

22.  (a)  What  is  lim  ? 

(b)  What  is     lim     Sm(x  +  y)? 

(x,y)^(0,0)      x  +  y 

(c)  What  is  lim 

(x,y)^(0,0)  xy 


2.2  |  Exercises  115 


23.  Examine  the  behavior  of  fix,  y)  =  x4y4/(x2  +  y4)3 
as  (x,  y)  approaches  (0,  0)  along  various  straight  lines. 
From  your  observations,  what  might  you  conjecture 
lim(XJ,)_i.(oio)  f(x,  y)  to  be?  Next,  consider  what  hap- 
pens when  (x,  y)  approaches  (0,  0)  along  the  curve 
x  =  y2.  Does  lim^-  yj^^o)  fix,  y)  exist?  Why  or  why 
not?' 

In  Exercises  24—27,  (a)  use  a  computer  to  graph  z  =  f(x,  y); 
(b)  use  your  graph  in  part  (a)  to  give  a  geometric  discussion 
as  to  whether  lim^  ^^^o)  f(x,  y)  exists;  (c)  give  an  analytic 
(i.e.,  nongraphical)  argument  for  your  answer  in  part  (b). 

4x2  +  2xy  +  5  y2 


^  24.  f(x,  y)  : 

^25.  f(x,y)-. 

^  26.  f{x,  y)  : 

27.  f(x,  y)  : 


3x2  +  5y2 


y 


x2  +  y2 

„5 


xy 


x2  +  v1 


x  sin  —     if  y  ^  0 


y 


o 


ify  =  0 


Some  limits  become  easier  to  identify  if  we  switch  to  a  different 
coordinate  system.  In  Exercises  28-33  switch  from  Cartesian  to 
polar  coordinates  to  evaluate  the  given  limits.  In  Exercises  34— 
37  switch  to  spherical  coordinates. 


28. 


29. 


30. 


31. 


32. 


33. 


34. 


35. 


36. 


37. 


lim 


lim 

*,v>^(0.0) 


lim 


lim     —  - 

z,}>)->-(0,0)  x2  +  y2 

x2 

lim     —  r 

x,y)-y(0,0)  X2  +  y2 


x2  +  xy  +  y2 
x2+y2 

x5  +  y4  -  3x3y  +  2x2  +  2y2 
x2  +  y2 


x  +  y 

lim 


lim 


2 

x  y 


x,j,z)^(0,o,o)  x2  +  y2  +  z2 
xyz 

lim  —  

x,y,z)^(0,0,0)  x2  +  y2  +  z2 


lim 


x2  +  y2 


*,y,z)-K0,0.0)  Jx2  +  y2  +z2 

XZ 

lim      —  =  - 

>,y,z)^-(0,0,0)  X2  +  V2  +  Z2 


In  Exercises  38—45,  determine  whether  the  functions  are  con- 
tinuous throughout  their  domains: 

38.  fix,y)  =  x2  +  2xy  -  y1 

39.  fix,  y,  z)  =  x2  +  3xyz  +  yz3  +2 


40.  g(x,y): 


x2  +  \ 


41.  h(x,  y)  =  cos 


x2-y2 
x2  +  \ 


42.  fix,  y)  =  cos2  x  —  2  sin2  xy 

if(x,y)/(0,0) 
0         if(x,y)  =  (0,0) 
x3  +  x2  +  xy2  +  y 


43.  fix,  y)  : 


x2-y2 
x^Ty2 


44.  g(x,y)  . 


x2  +  y2 
2 


45.  F(x,y,z)=  [xz  +  3xy 


if(x,y)/(0,  0) 
if(x,y)  =  (0,  0) 

xy 


2x2  +  y4  +  3  '      \v2  +  l 

46.  Determine  the  value  of  the  constant  c  so  that 

'  x3  +xy2  +2x2  +2y2 


g(x,  y)  = 


is  continuous. 


x  +  y2 
c 


if  ix,  y)  #  (0,  0) 
if  ix,  y)  =  (0,  0) 


47.  Show  that  the  function  /:  R3  — >  R  given  by  f(\)  = 
(2i  —  3j  +  k)  •  x  is  continuous. 

48.  Show  that  the  function  f:  R3  — >  R3  given  by  f(x)  = 
(6i  —  5k)  x  x  is  continuous. 

Exercises  49—53  involve  Definition  2.2  of  the  limit. 

49.  Consider  the  function  fix)  =  2x  —  3. 

(a)  Show  that  if  \x  -  5|  <  5,  then  |/(jc)  -  7|  <  25. 

(b)  Use  part  (a)  to  prove  that  limA-^5  fix)  =  7. 

50.  Consider  the  function  fix,  y)  =  2x  —  lOy  +  3. 

(a)  Show  that  if  ||(jc,  y)  -  (5,  1)||  <  S,  then  \x  -  5|  < 
S  and  |y  —  1 1  <  S. 

(b)  Use  part  (a)  to  show  that  if  ||(jc,  y)  -  (5,  1)||  <  S, 
then|/(A,y)-  3|  <  125. 

(c)  Show  that  lim(x  },)^(5i)  fix,  y)  =  3. 

51 .  If  A,  B,  and  C  are  constants  and  fix,  y)  =  Ax  +  By  + 
C,  show  that 

lim      fix,  y)  =  fix0,  y0)  =  Ax0  +  By0  +  C. 

(jc,y)-s-(x0,>'o) 


Chapter  2  |  Differentiation  in  Several  Variables 


52.  In  this  problem,  you  will  establish  rigorously  that 

lim     —          =  0. 

(x,y)M0,0)  x2  +  y2 

(a)  Show  that  \x\  <  \\{x,y)\\  and \y\  <  \\(x,y)\\. 

(b)  Showthat  |jc3  +  y3\  <  2(x2  +  y2fl2.  (Hint:  Begin 
with  the  triangle  inequality,  and  then  use  part  (a).) 

(c)  Show  that  if  0  <  ||(jc,  y)||  <  S,  then  \(x3  +  y3)/ 
(x2  +  y2)\  <  28. 

(d)  Now   prove   that   \im(x  v)^(0  o)(-*3  +  v3)/(x2  + 

y1)  =  o. 


53.  (a)  If  a  and  b  are  any  real  numbers,  show  that  2\ab\  < 
a2  +  b2. 

(b)  Let 

(x2-  y2\ 

Use  part  (a)  to  show  that  if  0  <  ||(jc,  y)\\  <  S,  then 
\f(x,y)\  <S2/2. 

(c)  Prove  that  lim(A: ,_V)_j.(o,o)  f(x,  y)  exists,  and  find  its 
value. 


2.3   The  Derivative 

Our  goal  for  this  section  is  to  define  the  derivative  of  a  function  f :  X  C  R"  — >  R" , 
where  n  and  m  are  arbitrary  positive  integers.  Predictably,  the  derivative  of  a 
vector-valued  function  of  several  variables  is  a  more  complicated  object  than  the 
derivative  of  a  scalar- valued  function  of  a  single  variable.  In  addition,  the  notion  of 
differentiability  is  quite  subtle  in  the  case  of  a  function  of  more  than  one  variable. 

We  first  define  the  basic  computational  tool  of  partial  derivatives.  After  do- 
ing so,  we  can  begin  to  understand  differentiability  via  the  geometry  of  tangent 
planes  to  surfaces.  Finally,  we  generalize  these  relatively  concrete  ideas  to  higher 
dimensions. 

Partial  Derivatives   

Recall  that  ifF:XcR^Risa  scalar- valued  function  of  one  variable,  then  the 
derivative  of  F  at  a  number  a  e  X  is 

F(a  +  h)-F(a) 
F(a)=hm  .  (1) 

fc->o  h 

Moreover,  F  is  said  to  be  differentiable  at  a  precisely  when  the  limit  in  equation 
(1)  exists. 


DEFINITION  3.1  Suppose  f:X  C  R"  R  is  a  scalar-  valued  function  of  n 
variables.  Let  x  =  (x\ ,  x%, . . . ,  xn)  denote  a  point  of  R" .  A  partial  function 
F  with  respect  to  the  variable  Xj  is  a  one-variable  function  obtained  from 
/  by  holding  all  variables  constant  except  x, .  That  is,  we  set  xj  equal  to  a 
constant  aj  for  j  ^  i.  Then  the  partial  function  in  x,  is  defined  by 

F(*i)  =        ,a2,...,Xi,...,  an). 


EXAMPLE  1  If/(x,  y)  =  (x2  -  y2)/(x2  +  y2),  then  the  partial  functions  with 
respect  to  x  are  given  by 


2  2 

X  —  ai 


F(x)  =  f(x,a2)  = 

x1  +  aj 


where  a2  may  be  any  constant.  If,  for  example,  a2  =  0,  then  the  partial  function  is 

x2 

F{x)  =  f(x,  0)  =  ^  =  1. 


2.3  |  The  Derivative  117 


y 


Domain  of/ 

/ 

Domain  of  F 

(restriction 

of/) 

Figure  2.46  The  function  /  of 
Example  1  is  defined  on  R2  - 
{(0,0)},  while  its  partial  function  F 
along  y  =  0  is  defined  on  the 
jc-axis  minus  the  origin. 


x 

Figure  2.47  Visualizing  the 
partial  derivative  f^(a,  b). 


z 


X 

Figure  2.48  Visualizing  the 
partial  derivative  §r(a,  b). 


Geometrically,  this  partial  function  is  nothing  more  than  the  restriction  of  /  to 
the  horizontal  line  y  =  0.  Note  that  since  the  origin  is  not  in  the  domain  of  /,  0 
should  not  be  taken  to  be  in  the  domain  of  F.  (See  Figure  2.46.)  ♦ 

Remark  In  practice,  we  usually  do  not  go  to  the  notational  trouble  of  explic- 
itly replacing  the  xj 's  (j  7^  i)  by  constants  when  working  with  partial  functions. 
Instead,  we  make  a  mental  note  that  the  partial  function  is  obtained  by  allowing 
only  one  variable  to  vary,  while  all  the  other  variables  are  held  fixed. 


DEFINITION  3.2   The  partial  derivative  of  /  with  respect  to  x;  is  the 

(ordinary)  derivative  of  the  partial  function  with  respect  to  x, .  That  is,  the 
partial  derivative  with  respect  to  x,  is  F'(x;),  in  the  notation  of  Definition 
3.1.  Standard  notations  for  the  partial  derivative  of  /  with  respect  to  x,-  are 

df 

~z    >     DXi  /(xi  ,...,*„),     and    fXi  (xu  x„). 
dx, 

Symbolically,  we  have 

9/      ,.    f(xi,...,xi  +  h,...,xn)-f(xi,...,xn) 

—  =  hm  .  (2) 

dxi    h^o  h 


By  definition,  the  partial  derivative  is  the  (instantaneous)  rate  of  change  of  / 
when  all  variables,  except  the  specified  one,  are  held  fixed.  In  the  case  where  / 
is  a  (scalar-valued)  function  of  two  variables,  we  can  understand 

ox 

geometrically  as  the  slope  at  the  point  (a,  b,  f(a,  b))  of  the  curve  obtained  by 
intersecting  the  surface  z  =  fix,  y)  with  the  plane  y  =  b,  as  shown  in  Figure  2.47. 
Similarly, 

-^-(a,  b) 
dy 

is  the  slope  at  (a,  b,  f(a,  fr))ofthe  curve  formed  by  the  intersection  of  z  =  f(x,  y) 
and  x  =  a,  shown  in  Figure  2.48. 

EXAMPLE  2  For  the  most  part,  partial  derivatives  are  quite  easy  to  compute, 
once  you  become  adept  at  treating  variables  like  constants.  If 

f(x,  y)  =  x2y  +  cos(x  +  y), 

then  we  have 

—  =  2xy  -  sin(x  +  y). 
dx 

(Imagine  y  to  be  a  constant  throughout  the  differentiation  process.)  Also 

df  7 

—  =  x  —  sin(x  +  y). 

dy 


Chapter  2  |  Differentiation  in  Several  Variables 


Figure  2.49  The  tangent  line  to 
y  =  F(x)  at  x  =  a  has  equation 
y  =  F(a)  +  F'(a)(x  -  a). 


(Imagine  x  to  be  a  constant.)  Similarly,  if  g(x,  y)  =  xy/(x2 
quotient  rule  of  ordinary  calculus,  we  have 

(x2  +  y2)y  -  xy(2x)      y(y2  - 
gx(x,  y)  =  - 


y2),  then,  from  the 


x2) 


and 


?y(*>  y  )  = 


(x2  +  y2)2 
(x2  +  y2)x  —  xy(2y) 


(x2  +  y2)2  ' 
x(x2  —  y2) 


(x2  +  y2)2  (x2  +  y2)2  ' 

Note  that,  of  course,  neither  g  nor  its  partial  derivatives  are  defined  at  (0, 0).  ♦ 

EXAMPLE  3  Occasionally,  it  is  necessary  to  appeal  explicitly  to  limits  to  eval- 
uate partial  derivatives.  Suppose  /:  R2  —>  R  is  defined  by 


/(*.  y) 


3x  y  —  y 


x2  +  y2 


if  (x,  y)  #  (0,  0) 
if(x,y)  =  (0,0) 


Then,  for  (x,  y)  ^  (0,  0),  we  have 
9/  8xy3 


3x      (x2  +  y2) 


2\2 


and 


df 
dy 


3x  —  6x  y 


2, ,2 


(x2  +  y2)2 


3/  3/ 

But  what  should  — (0,  0)  and  — (0,  0)  be?  To  find  out,  we  return  to  Definition 

dx  dy 

3.2  of  the  partial  derivatives: 


and 


df 
dy 


df 
dx 


(0,  0) 


(0,0) 


lim 


/(0  +  fc.O) 


 =  hm  

h-+o  h 


lim 

h^0 


f(0,  0  +  h)-  /(0,  0) 


lim 


lim  -1  = 


Tangency  and  Differentiability  

If  F:  X  c  R  ^  R  is  a  scalar- valued  function  of  one  variable,  then  to  have  F 
differentiable  at  a  number  a  e  X  means  precisely  that  the  graph  of  the  curve 
y  =  F(x)  has  a  tangent  line  at  the  point  (a,  F(a)).  (See  Figure  2.49.)  Moreover, 
this  tangent  line  is  given  by  the  equation 


y  =  F(a)  +  F'(a)(x  -  a). 


(3) 


If  we  define  the  function  H(x)  to  be  F(a)  +  F'(a)(x  —  a)  (i.e.,  H(x)  is  the  right 
side  of  equation  (3)  that  gives  the  equation  for  the  tangent  line),  then  H  has  two 
properties: 

1.  H(a)  =  F(a) 

2.  H'(a)  =  F'(a). 

In  other  words,  the  line  defined  by  y  =  H(x)  passes  through  the  point  (a,  F(a)) 
and  has  the  same  slope  at  (a,  F(a))  as  the  curve  defined  by  y  =  F(x).  (Hence, 
the  term  "tangent  line.") 

Now  suppose  /:lcR2^Risa  scalar-valued  function  of  two  variables, 
where  X  is  open  in  R2.  Then  the  graph  of  /  is  a  surface.  What  should  the  tangent 
plane  to  the  graph  of  z  =  f(x,  y)  at  the  point  (a,  b,  f(a,  b))  be?  Geometrically, 


2.3  |  The  Derivative  119 


(a,b,f{a,b)) 


Figure  2.50  The  plane  tangent 
toz  =  f(x,  y)  at 
(a,b,f(a,b)). 


y  =  b 


Figure  2.51  The  tangent  plane  at 
(a,  b,  f(a,  b))  contains  the  lines 
tangent  to  the  curves  formed  by 
intersecting  the  surface 
z  =  f(x,  y)  by  the  planes  x  =  a 
and  y  =  b. 


the  situation  is  as  depicted  in  Figure  2.50.  From  our  earlier  observations,  we 
know  that  the  partial  derivative  fx(a,  b)  is  the  slope  of  the  line  tangent  at  the 
point  (a,  b,  f(a,  b))  to  the  curve  obtained  by  intersecting  the  surface  z  =  f(x,  y) 
with  the  plane  y  =  b.  (See  Figure  2.51.)  This  means  that  if  we  travel  along  this 
tangent  line,  then  for  every  unit  change  in  the  positive  x -direction,  there's  a  change 
of  fx  (a ,  b)  units  in  the  z-direction.  Hence,  by  using  formula  ( 1)  of  §  1 .2,  the  tangent 
line  is  given  in  vector  parametric  form  as 

li(0  =  (a,  b,  /(a,  b))  +      0,  fx(a,  b)). 
Thus,  a  vector  parallel  to  this  tangent  line  is 

u  =  i  +  fx(a,b)k. 

Similarly,  the  partial  derivative  fy(a,  b)  is  the  slope  of  the  line  tangent  at  the  point 
(a,  b,  f(a,  b))  to  the  curve  obtained  by  intersecting  the  surface  z  =  f(x,  y)  with 
the  plane  x  —  a.  (Again  see  Figure  2.51.)  Consequently,  the  tangent  line  is  given 
by 

12(0  =  (a,  b,  f(a,  b))  +  t(0,  1,  fy(a,  b)), 
so  a  vector  parallel  to  this  tangent  line  is 

v  =  j  +  fy(a,b)k. 

Both  of  the  aforementioned  tangent  lines  must  be  contained  in  the  plane  tangent 
to  z  =  f{x,  y)  at  (a,  b,  f(a,  b)),  if  one  exists.  Hence,  a  vector  n  normal  to  the 
tangent  plane  must  be  perpendicular  to  both  u  and  v.  Therefore,  we  may  take  n 
to  be 

n  =  u  x  v  =  -fx(a,  b)\-  fy(a,  b)\  +  k. 

Now,  use  equation  ( 1 )  of  §  1 .5  to  find  that  the  equation  for  the  tangent  plane — that 
is,  the  plane  through  (a,  b,  f(a,  b))  with  normal  n — is 

(-fx(a,  b),  -fy(a,  b),  1)  •  (x  -  a,  y  -  b,  z  -  f(a,  bj)  =  0 

or,  equivalently, 

-fx(a,  b)(x  -a)-  fy(a,  b)(y  -b)  +  z-  f(a,  b)  =  0. 

By  rewriting  this  last  equation,  we  have  shown  the  following  result: 


THEOREM  3.3  If  the  graph  of  z  =  f(x,  y)  has  a  tangent  plane  at 
(a,  b,  f(a,  b)),  then  that  tangent  plane  has  equation 


z  =  f(a,  b)  +  fx(a,  b)(x  -  a)  +  fy(a,  b)(y  -  b). 


(4) 


Note  that  if  we  define  the  function  h(x,  y)  to  be  equal  to  f(a,  b)  +  fx(a,  b)(x  — 
a)  +  fy(a,  b)(y  —  b)  (i.e.,  h(x,  y)  is  the  right  side  of  equation  (4)),  then  h  has  the 
following  properties: 


1.  h(a,b)  =  f(a,b) 
dh  df 
ox  ax 


dh  df 
and    — (a,b)= — (a,  b). 
dy  dy 


In  other  words,  h  and  its  partial  derivatives  agree  with  those  of  /  at  (a,  b). 

It  is  tempting  to  think  that  the  surface  z  =  fix,  y)  has  a  tangent  plane  at 
(a,  b,  f(a,  b))  as  long  as  you  can  make  sense  of  equation  (4),  that  is,  as  long  as  the 


1  20       Chapter  2  |  Differentiation  in  Several  Variables 


Figure  2.52  If  two  points 
approach  (0,  0,  0)  while  remaining 
on  one  face  of  the  surface 
described  in  Example  4,  the 
limiting  plane  they  and  (0,  0,  0) 
determine  is  different  from  the  one 
determined  by  letting  the  two 
points  approach  (0,  0,  0)  while 
remaining  on  another  face. 


partial  derivatives  fx(a,  b)  and  fy(a,  b)  exist.  Indeed  this  would  be  analogous  to 
the  one-variable  situation  where  the  existence  of  the  derivative  and  the  existence 
of  the  tangent  line  mean  exactly  the  same  thing.  However,  it  is  possible  for  a 
function  of  two  variables  to  have  well-defined  partial  derivatives  (so  that  equation 
(4)  makes  sense)  yet  not  have  a  tangent  plane. 

EXAMPLE  4  Let  f(x,  y)  =  \\x\  -  \y\\  -  \x\  -  |y|  and  consider  the  surface 
defined  by  the  graph  of  z  =  f(x,  y)  shown  in  Figure  2.52.  The  partial  derivatives 
of  /  at  the  origin  may  be  calculated  from  Definition  3.2  as 


A(0,  0)  =  lim 

h— >0 


/(0  +  ft,0)-/(0,  0) 


=  lim 


and 


fy(0,  0)  =  lim 

h^0 


/(O,  0  +  h)-  /(O,  0) 


=  lim 

h^0 


h 


=  lim  0  =  0 


=  lim  0  =  0. 

h^0 


(Indeed  the  partial  functions  F(x)  =  f(x,  0)  and  G(y)  =  f(0,  y)  are  both  identi- 
cally zero  and  thus,  have  zero  derivatives.)  Consequently,  if  'the  surface  in  question 
has  a  tangent  plane  at  the  origin,  then  equation  (4)  tells  us  that  it  has  equation 
z  =  0.  But  there  is  no  geometric  sense  in  which  the  surface  z  =  f(x,  y)  has  a 
tangent  plane  at  the  origin.  If  we  think  of  a  tangent  plane  as  the  geometric  limit  of 
planes  that  pass  through  the  point  of  tangency  and  two  other  "moving"  points  on 
the  surface  as  those  two  points  approach  the  point  of  tangency,  then  Figure  2.52 
shows  that  there  is  no  uniquely  determined  limiting  plane.  ♦ 


Example  4  shows  that  the  existence  of  a  tangent  plane  to  the  graph  of 
z  =  f(x,  y)  is  a  stronger  condition  than  the  existence  of  partial  derivatives.  It 
turns  out  that  such  a  stronger  condition  is  more  useful  in  that  theorems  from  the 
calculus  of  functions  of  a  single  variable  carry  over  to  the  context  of  functions 
of  several  variables.  What  we  must  do  now  is  find  a  suitable  analytic  definition  of 
differentiability  that  captures  this  idea.  We  begin  by  looking  at  the  definition  of 
the  one-variable  derivative  with  fresh  eyes. 

By  replacing  the  quantity  a  +  h  by  the  variable  x ,  the  limit  equation  in  formula 
(1)  may  be  rewritten  as 

F(x)  -  F(a) 
F  (a)  =  hm  . 


This  is  equivalent  to  the  equation 

F(x) 


lim 


F(a) 


F'(a)  =  0. 


The  quantity  F'(a)  does  not  depend  on  x  and  therefore  may  be  brought  inside  the 
limit.  We  thus  obtain  the  equation 

F(x)  -  F(a) 


lim 


F'(a)  =0. 


Finally,  some  easy  algebra  enables  us  to  conclude  that  the  function  F  is  differen- 
tiable  at  a  if  there  is  a  number  F'(a)  such  that 


lim 


F(x)  -  [F(a)  +  F'(a)(x  -  a)] 


=  0. 


(5) 


What  have  we  learned  from  writing  equation  (5)?  Note  that  the  expression  in 
brackets  in  the  numerator  of  the  limit  expression  in  equation  (5)  is  the  function 


2.3  |  The  Derivative  121 


(x,F(x)) 

^\(a,  F(a))  \ 

(x,  H(x)^ 

a  x 

Figure  2.53  If  F  is  differentiable 
at  a,  the  vertical  distance  between 
F(x)  and  H(x)  must  approach 
zero  faster  than  the  horizontal 
distance  between  x  and  a  does. 


H(x)  that  was  used  to  define  the  tangent  line  to  y 
may  rewrite  equation  (5)  as 


F(x)  at  (a,  F(a)).  Thus,  we 


lim 


F{x)  -  H(x) 


=  0. 


For  the  limit  above  to  be  zero,  we  certainly  must  have  that  the  limit  of  the  numerator 
is  zero.  But  since  the  limit  of  the  denominator  is  also  zero,  we  can  say  even  more, 
namely,  that  the  difference  between  the  y-values  of  the  graph  of  F  and  of  its  tangent 
line  must  approach  zero  faster  than  x  approaches  a.  This  is  what  is  meant  when 
we  say  that  "H  is  a  good  linear  approximation  to  F  near  a."  (See  Figure  2.53.) 
Geometrically,  it  means  that,  near  the  point  of  tangency,  the  graph  of  y  =  F(x) 
is  approximately  straight  like  the  graph  of  y  =  H(x). 

If  we  now  pass  to  the  case  of  a  scalar- valued  function  fix ,  y)  of  two  variables, 
then  to  say  that  z  =  fix,  y)  has  a  tangent  plane  at  (a,  b,  f(a,  b))  (i.e.,  that  /  is 
differentiable  at  ia,  b))  should  mean  that  the  vertical  distance  between  the  graph 
of  /  and  the  "candidate"  tangent  plane  given  by 

z  =  h(x,  y)  =  fia,  b)  +  fx(a,  b){x  -a)+  fy(a,  b\y  -  b) 

must  approach  zero  faster  than  the  point  (jc,  y)  approaches  (a,b).  (See  Fig- 
ure 2.54.)  In  other  words,  near  the  point  of  tangency,  the  graph  of  z  =  fix,  y)  is 
approximately  flat  just  like  the  graph  of  z  =  h(x,  y).  We  can  capture  this  geometric 
idea  with  the  following  formal  definition  of  differentiability: 


DEFINITION  3.4  Let  X  be  open  in  R2  and  /:  X  C  R2  ->  R  be  a  scalar- 
valued  function  of  two  variables.  We  say  that  /  is  differentiable  at  (a,  b)  e  X 
if  the  partial  derivatives  fx(a,  b)  and  fy(a,  b)  exist  and  if  the  function 


h(x<  y)  =  f(a<  b)  +  fx(fl,  b\x  -a)  +  fy(a,  b)(y 
is  a  good  linear  approximation  to  /  near  (a,  b) — that  is,  if 

fix,y)-h(x,y) 


b) 


lim 


ix,y)-ia,b)\\ 


=  0. 


Moreover,  if  /  is  differentiable  at  (a,  b),  then  the  equation  z  =  h(x,  y)  de- 
fines the  tangent  plane  to  the  graph  of  /  at  the  point  (a,  b,  fia,  b)).  If  / 
is  differentiable  at  all  points  of  its  domain,  then  we  simply  say  that  /  is 
differentiable. 


(x,y,f(x,y)) 

(x,y,h(x,y)) 
ia,b,fia,b)) 


Figure  2.54  If  /  is  differentiable  at  (a,  b),  the  distance 
between  fix,  y)  and  h(x,  y)  must  approach  zero  faster  than 
the  distance  between  (x,  y)  and  (a,  b)  does. 


Chapter  2  |  Differentiation  in  Several  Variables 


EXAMPLE  5  Let  us  return  to  the  function  f(x,  y)  =  \\x\  -  \y\\  -  \x\  -  \y\  of 
Example  4.  We  already  know  that  the  partial  derivatives  fx  (0,  0)  and  /v(0,  0) 
exist  and  equal  zero.  Thus,  the  function  h  of  Definition  3.4  is  the  zero  function. 
Consequently,  /  will  be  differentiable  at  (0,0)  just  in  case 

f(x,y)-h(x,y)  f(x,y) 

hm   =  hm   

(a-,y)^(0,0)  ||(*,  y)  -  (0,  0)||      C*,y)->-(0,0)  ||(x,  y)|| 

\x\-\y\\-\x\-\y\ 


=  lim 


(x,v)^(o,o)      yx2  + 


r 


is  zero.  However,  it  is  not  hard  to  see  that  the  limit  in  question  fails  to  exist.  Along 
the  line  v  =  0,  we  have 

f(x,y)      IM-01- W-I0I       0  Q 


ll(*,y)ll  \x 
but  along  the  line  y  =  x,  we  have 

f(x,y)  =  lk|-|x||-|x|-|x|  =  -2|*|  =  _^ 
\\(x,y)\\  s/x2  +  x2  </2\x\ 

Hence,  /  fails  to  be  differentiable  at  (0,  0)  and  has  no  tangent  plane  at  (0,  0,  0)> 

The  limit  condition  in  Definition  3.4  can  be  difficult  to  apply  in  practice. 
Fortunately,  the  following  result,  which  we  will  not  prove,  simplifies  matters  in 
many  instances.  Recall  from  Definition  2.3  that  the  phrase  "a  neighborhood  of 
a  point  P  in  a  set  X"  just  means  an  open  set  containing  P  and  contained  in  X. 

THEOREM  3.5  Suppose  X  is  open  in  R2.  If  /:  X  —>  R  has  continuous  partial 
derivatives  in  a  neighborhood  of  (a,  b)  in  X,  then  /  is  differentiable  at  (a,  b). 

A  proof  of  a  more  general  result  (Theorem  3. 10)  is  provided  in  the  addendum 
to  this  section. 

EXAMPLE  6  Let  /(*,  y)  =  x2  +  2y2.  Then  df/dx  =  2x  and  df/dy  =  Ay, 
both  of  which  are  continuous  functions  on  all  of  R2.  Thus,  Theorem  3.5  implies 
that  /  is  differentiable  everywhere.  The  surface  z  =  x2  +  2y2  must  therefore 
have  a  tangent  plane  at  every  point.  At  the  point  (2,  —  1),  for  example,  this  tangent 
plane  is  given  by  the  equation 

z  =  6  +  4(x  -  2)  -  4(y  +  1) 

(or,  equivalently,  by  Ax  —  Ay  —  z  =  6).  ♦ 

While  we're  on  the  subject  of  continuity  and  differentiability,  the  next  result  is 
the  multivariable  analogue  of  a  familiar  theorem  about  functions  of  one  variable. 

THEOREM  3.6  If  /:  X  C  R2  R  is  differentiable  at  (a,  b),  then  it  is  continu- 
ous at  (a,  b). 


2.3  |  The  Derivative  123 


EXAMPLE  7   Let  the  function  /:  R2 

/(*.  y)  -- 


2  2 

x  y 


R  be  denned  by 

if(jc,  30^(0,0) 


x4  +  y4 

0        if(x,y)  =  (0,0) 

The  function  /  is  not  continuous  at  the  origin,  since  lim^  ^j^^o.o)  fixi  y)  does 
not  exist.  (However,  /  is  continuous  everywhere  else  in  R  .)  By  Theorem  3.6,  / 
therefore  cannot  be  differentiate  at  the  origin.  Nonetheless,  the  partial  derivatives 
of  /  do  exist  at  the  origin,  and  we  have 


and 


fix,  0)  = 


7(0,  .v)  = 


o 


x4  +  0 


0+ v4 


0 


0 


df 

dx 


df 
dy 


(0,0)  =  0, 


(0,0)  =  0, 


since  the  partial  functions  are  constant.  Thus,  we  see  that  if  we  want  something 
like  Theorem  3.6  to  be  true,  the  existence  of  partial  derivatives  alone  is  not 
enough.  ♦ 


Differentiability  in  General  

It  is  not  difficult  now  to  see  how  to  generalize  Definition  3.4  to  three  (or  more) 
variables:  For  a  scalar-valued  function  of  three  variables  to  be  differentiable  at  a 
point  (a,  b,  c),  we  must  have  that  (i)  the  three  partial  derivatives  exist  at  (a,  b,  c) 
and  (ii)  the  function  h :  R3  — >  R  defined  by 

h(x,  y,  z)  =  f(a,  b,  c)  +  fx{a,  b,  c)(x  —  a) 

+  fy(a,  b,  c)(y  -b)  +  ft(a,  b,  c)(z  -  c) 

is  a  good  linear  approximation  to  /  near  (a,  b,  c).  In  other  words,  (ii)  means  that 

f(x,y,z)-h(x,y,z) 
hm   =  0. 

(x,y,zy+(a,b,c)  ||(x,  V,  z)  ~  (a,  b,  C)\\ 

The  passage  from  three  variables  to  arbitrarily  many  is  now  straightforward. 


DEFINITION  3.7  Let  X  be  open  in  R"  and  /:  X  R  be  a  scalar-valued 
function;  let  a  =  (a\,  a%,  . . . ,  a„)  e  X.  We  say  that  /  is  differentiable  at 
a  if  all  the  partial  derivatives  fXj(a),  i  =  1, . . . ,  n,  exist  and  if  the  function 
h:Rn  ->  R  defined  by 

H*)  =  /(a)  +  Ai(a)(xi  -  fli)  +  -  a2) 

H  h  fXn{&)(xn  -  an)  (6) 

is  a  good  linear  approximation  to  /  near  a,  meaning  that 

hm  =  0. 

*->■»    II  x  —  all 


1  24       Chapter  2  |  Differentiation  in  Several  Variables 


We  can  use  vector  and  matrix  notation  to  rewrite  things  a  bit.  Define  the 
gradient  of  a  scalar- valued  function  /:  X  C  R"      R  to  be  the  vector 

df_  df_  df_ 


V/(x)  = 
Consequently, 


V/(a)  =  (A-1(a),/X2(a),...,  A„(a)). 

Alternatively,  we  can  use  matrix  notation  and  define  the  derivative  of  /  at  a, 
denoted  Df(a),  to  be  the  row  matrix  whose  entries  are  the  components  of  V /(a); 
that  is, 

»/(»)  =  [/*(»)  M»)  ■■■  /,„(a)]. 

Then,  by  identifying  the  vector  x  —  a  with  the  n  x  1  column  matrix  whose  entries 
are  the  components  of  x  —  a,  we  have 


V/(a).(x-a)  =  D/(a)(x-a)  =  [/Tl(a)  /,2(a)  •••  A„(a)] 


X\  —  d\ 

x2  -  a2 


=  /xi(a)(*i  -  fli)  +  /x2(a)(x2  -  a2) 

H  1-  /v„(a)(x„  -  an). 

Hence,  vector  notation  allows  us  to  rewrite  equation  (6)  quite  compactly  as 
ft(x)=/(a)  +  V/(a).(x-a). 

Thus,  to  say  that  /z  is  a  good  linear  approximation  to  /  near  a  in  equation  (6) 
means  that 

||x  -  a|| 

Compare  equation  (7)  with  equation  (5).  Differentiability  of  functions  of  one  and 
several  variables  should  really  look  very  much  the  same  to  you.  It  is  worth  noting 
that  the  analogues  of  Theorems  3.5  and  3.6  hold  in  the  case  of  n  variables. 

The  gradient  of  a  function  is  an  extremely  important  construction,  and  we 
consider  it  in  greater  detail  in  §2.6. 

You  may  be  wondering  what,  if  any,  geometry  is  embedded  in  this  general 
notion  of  differentiability.  Recall  that  the  graph  of  the  function  /:lcR"->R 
is  the  hypersurface  in  R"+1  given  by  the  equation  xn+\  =  f(xi,  x2,  ...  ,  xn). 
(See  equation  (2)  of  §2.1.)  If  /  is  differentiable  at  a,  then  the  hypersurface  deter- 
mined by  the  graph  has  a  tangent  hyperplane  at  (a,  /(a))  given  by  the  equation 

x„+i  =  h(x{  ,x2,...,x„)  =  /(a)  +  V/(a)  •  (x  -  a) 

=  /(a)  +  £>/(a)(x  -  a).  (8) 

Compare  equation  (8)  with  equation  (3)  for  the  tangent  line  to  the  curve  y  =  F(x) 
at  (a,  F(a)).  Although  we  cannot  visualize  the  graph  of  a  function  of  more  than 
two  variables,  nonetheless,  we  can  use  vector  notation  to  lend  real  meaning  to 
tangency  in  n  dimensions. 


EXAMPLE  8  Before  we  drown  in  a  sea  of  abstraction  and  generalization,  let's 
do  some  concrete  computation.  An  example  of  an  "« -dimensional  paraboloid"  in 


2.3  |  The  Derivative       1 25 


Rn+1  is  given  by  the  equation 


Xn+\  —  X\  +  *2  +  ■  ■  ■  +  X2, 


that  is,  by  the  graph  of  the  function /(xi,  . . .  ,x„)  =  x2  +  x|  H  +  x2.  We  have 

—  =  2x;- ,     z  =  1,  2, . . . ,  n, 

dxi 

so  that 


V/(*i 


x„)  =  (2xi,  2x2, 


2x„). 


Note  that  the  partial  derivatives  of  /  are  continuous  everywhere.  Hence,  the 
n -dimensional  version  of  Theorem  3.5  tells  us  that  /  is  differentiate  everywhere. 
In  particular,  /  is  differentiable  at  the  point  (1,2,  ...,«), 


V/(l,2,  ...,n)  =  (2,4,  ...,2n), 


and 


Df(l,2,...,n)=  [2  4   ■■■  In]. 
Thus,  the  paraboloid  has  a  tangent  hyperplane  at  the  point 


(1,2, \2  +  22  +  ---+n2) 


whose  equation  is  given  by  equation  (8): 


x,1+1  =(l2  +  22  + 


=  (l2  +  22  + 


+  n2)+[2  4   ■■■  2n] 


X\  —  1 
x2  -  2 


+  n2)  +  2(xi  -  1)  +  4(x2  -!)  +  ■■■+  2n{xn  -  n) 


=  (l2  +  22  + 


+  n2)  +  2x]  +  4x2  +  • 
(2  ■  1  +  4  ■  2  H  \-2n-n) 


=  2xi  +  4x2  H  h  2«x„  -  (l2  +  22  + 


2«x„ 


+  «2) 


=  2i'x, 


n(n  +  l)(2n  +  1) 


(The  formula  l2  +  22  +  ■  ■  ■  +  n2  =  n(n  +  l)(2n  +  l)/6  is  a  welhlmown  identity, 
encountered  when  you  first  learned  about  the  definite  integral.  It's  straightforward 
to  prove  using  mathematical  induction.)  ♦ 

At  last  we're  ready  to  take  a  look  at  differentiability  in  the  most  general  setting 
of  all.  Let  X  be  open  in  R"  and  let  f:  X  —>  Rm  be  a  vector- valued  function  of  n 
variables.  We  define  the  matrix  of  partial  derivatives  of  f,  denoted  Df,  to  be 


Chapter  2  |  Differentiation  in  Several  Variables 


the  m  x  n  matrix  whose  ijth  entry  is  df/dxj,  where  f:X  C  R" 
component  function  of  f.  That  is, 


R  is  the  (th 


Df(x\,x2,  ...,xn)  = 


dxi 
dx\ 


dxi 


dX2 

d_h 

dx2 


Mm 
dX2 


dxn 

M. 

dx„ 


Mm 

dxn 


The  ith  row  of  Df  is  nothing  more  than  Dfj — and  the  entries  of  Df  are  precisely 
the  components  of  the  gradient  vector  V  f,  (Indeed  in  the  case  where  m  =  1, 
V/  and  Df  mean  exactly  the  same  thing.) 

EXAMPLE  9  Suppose  f:  R3  ->  R2  is  given  by  f(x,  y,  z)  =  (x  cos  y  +  z,  xj). 
Then  we  have 


Df(x,y,z)  = 


cos  y   —x  sin  y 

y  ^ 


We  generalize  equation  (7)  and  Definition  3.7  in  an  obvious  way  to  make  the 
following  definition: 


DEFINITION  3.8  (Grand  definition  of  differentiability)  Let  X  c 
R"  be  open,  let  f :  X  ->•  R",  and  let  a  e  X.  We  say  that  f  is  differentiable  at 
a  if  Df(a)  exists  and  if  the  function  h:  R"  ->  R"  defined  by 

h(x)  =  f(a)  +  Df(a)(x  -  a) 

is  a  good  linear  approximation  to  f  near  a.  That  is,  we  must  have 


lim 


II fW  -  h(x) 


lim 


|f(x)-[f(a)  +  Df(a)(x-a) 


=  0. 


Some  remarks  are  in  order.  First,  the  reason  for  having  the  vector  length 
appearing  in  the  numerator  in  the  limit  equation  in  Definition  3.8  is  so  that  there 
is  a  quotient  of  real  numbers  of  which  we  can  take  a  limit.  (Definition  3 . 7  concerns 
scalar-valued  functions  only,  so  there  is  automatically  a  quotient  of  real  numbers.) 
Second  the  term  Df(a)(x  —  a)  in  the  definition  of  h  should  be  interpreted  as  the 
product  of  the  m  x  n  matrix  Df(a)  and  the  nxl  column  matrix 


x2 


Q2 


Because  of  the  consistency  of  our  definitions,  the  following  results  should 
not  surprise  you: 


2.3  |  The  Derivative       1 27 


THEOREM  3.9  If  f:  X  C  R"  Rm  is  differentiable  at  a,  then  it  is  continuous 
at  a. 


THEOREM  3.10  If  f:lcR"^  R"'  is  such  that,  for  i  =  l,...,m  and 
j  =  1, . . . ,  n,  all  dfi/dxj  exist  and  are  continuous  in  a  neighborhood  of  a  in 
X,  then  f  is  differentiable  at  a. 


THEOREM  3.1 1  A  function  f:  X  C  R"  ->  Rm  is  differentiable  ataeX  (in  the 
sense  of  Definition  3.8)  if  and  only  if  each  of  its  component  functions  fi'.X  C 
R"  — »  R,  i'  =  1, . . . ,  m,  is  differentiable  at  a  (in  the  sense  of  Definition  3.7). 

The  proofs  of  Theorems  3.9,  3.10,  and  3.11  are  provided  in  the  addendum 
to  this  section.  Note  that  Theorems  3.10  and  3.1 1  frequently  make  it  a  straight- 
forward matter  to  check  that  a  function  is  differentiable:  Just  look  at  the  partial 
derivatives  of  the  component  functions  and  verify  that  they  are  continuous.  Thus, 
in  many — but  not  all — circumstances,  we  can  avoid  working  directly  with  the 
limit  in  Definition  3.8. 


EXAMPLE  10   The  function  g:  R3  -  {(0,  0,  0)}      R3  given  by 

3 


g(*>  y,z) 


x2  +  y2  +  z2 


has 


Dg(x,  y,  z)  = 


-6z 


—6x  —6y 

(x2  +  y2  +  z2)2  (x2  +  y2  +  z2)2  (x2  +  y2  +  z2)2 
y                        x  0 

z  0  x 


Each  of  the  entries  of  this  matrix  is  continuous  over  R3  —  {(0,  0,  0)}.  Hence,  by 
Theorem  3. 10,  g  is  differentiable  over  its  entire  domain.  ♦ 


What  Is  a  Derivative?  

Although  we  have  defined  quite  carefully  what  it  means  for  a  function  to  be 
differentiable,  the  derivative  itself  has  really  taken  a  "backseat"  in  the  preceding 
discussion.  It  is  time  to  get  some  perspective  on  the  concept  of  the  derivative. 

In  the  case  of  a  (differentiable)  scalar-valued  function  of  a  single  variable, 
/:XcR->R,  the  derivative  f'ia)  is  simply  a  real  number,  the  slope  of  the 
tangent  line  to  the  graph  of  /  at  the  point  (a,  f(a)).  From  a  more  sophisticated 
(and  slightly  less  geometric)  point  of  view,  the  derivative  f'{a)  is  the  number  such 
that  the  function 

h(x)  =  /(c)  +  f'(a)(x  -  a) 

is  a  good  linear  approximation  to  f(x)  for  x  near  a.  (And,  of  course,  y  =  h(x)  is 
the  equation  of  the  tangent  line.) 

If  a  function  /:XcR"-^Rof«  variables  is  differentiable,  there  must 
exist  n  partial  derivatives  3  f/dx\ , . . . ,  df/dx„ .  These  partial  derivatives  form  the 
components  of  the  gradient  vector  V/  (or  the  entries  of  the  1  x  n  matrix  Df).  It 


Chapter  2  |  Differentiation  in  Several  Variables 


is  the  gradient  that  should  properly  be  considered  to  be  the  derivative  of  /,  but  in 
the  following  sense:  V/(a)  is  the  vector  such  that  the  function  h:  R"  —>  R  given 
by 

fc(x)=/(a)  +  V/(a).(X-a) 

is  a  good  linear  approximation  to  /(x)  for  x  near  a.  Finally,  the  derivative  of  a 
differentiable  vector- valued  function  f :  X  c  R"  — >  R"!  may  be  taken  to  be  the 
matrix  Df  of  partial  derivatives,  but  in  the  sense  that  the  function  h:  R"  — >  R"1 
given  by 

h(x)  =  f(a)  +  Df(a)(x  -  a) 

is  a  good  linear  approximation  to  f(x)  near  a.  You  should  view  the  derivative 
Df(a)  not  as  a  "static"  matrix  of  numbers,  but  rather  as  a  matrix  that  defines  a 
linear  mapping  from  R"  to  R'".  (See  Example  5  of  §1.6.)  This  is  embodied  in 
the  limit  equation  of  Definition  3.8  and,  though  a  subtle  idea,  is  truly  the  heart  of 
differential  calculus  of  several  variables. 

In  fact,  we  could  have  approached  our  discussion  of  differentiability  much 
more  abstractly  right  from  the  beginning.  We  could  have  defined  a  function  f:lc 
R"  — >  R°'  to  be  differentiable  at  a  point  a  6  X  to  mean  that  there  exists  some 
linear  mapping  L:  R"  — >  R'"  such  that 

Hm  ||f(x)-[f(a)  +  L(x-a)]||  =  Q 
x^a  ||x  —  a|| 

Recall  that  any  linear  mapping  L:  R"  — >  R'"  is  really  nothing  more  than  multipli- 
cation by  a  suitable  m  x  n  matrix  A  (i.e.,  that  L(y)  =  Ay).  It  is  possible  to  show 
that  if  there  is  a  linear  mapping  that  satisfies  the  aforementioned  limit  equation, 
then  the  matrix  A  that  defines  it  is  both  uniquely  determined  and  is  precisely  the 
matrix  of  partial  derivatives  Df(a).  (See  Exercises  60-62  where  these  facts  are 
proved.)  However,  to  begin  with  such  a  definition,  though  equivalent  to  Definition 
3.8,  strikes  us  as  less  well  motivated  than  the  approach  we  have  taken.  Hence,  we 
have  presented  the  notions  of  differentiability  and  the  derivative  from  what  we 
hope  is  a  somewhat  more  concrete  and  geometric  perspective. 


Addendum:  Proofs  of  Theorems  3.9,  3.10,  and  3.11   

Proof  of  Theorem  3.9  We  begin  by  claiming  the  following:  Let  xeR"  and 
B  =  (bjj)  be  an  m  x  n  matrix.  If  y  =  Bx,  (so  y  6  Rm),  then 

llyll  <TOI,  (9) 

/  ,\l/2 

where  K  =  (  >_\  .  bf- 1  .  We  postpone  the  proof  of  (9)  until  we  establish  the 
main  theorem. 

To  show  that  f  is  continuous  at  a,  we  will  show  that  ||f(x)  —  f(a)||  —>  0  as 
x  ->  a.  We  do  so  by  using  the  fact  that  f  is  differentiable  at  a  (Definition  3.8). 
We  have 

||f(x)  -  f(a)||  =  ||f(x)  -  f(a)  -  Df(a)(x  -  a)  +  Df(a)(x  -  a)|| 

<  ||f(x)  -  f(a)  -  £>f(a)(x  -  a)||  +  ||Df(a)(x  -  a)||,  (10) 

using  the  triangle  inequality.  Note  that  the  first  term  in  the  right  side  of  inequality 
(10)  is  the  numerator  of  the  limit  expression  in  Definition  3.8.  Thus,  since  f  is 


2.3  |  The  Derivative       1 29 


differentiable  at  a,  we  can  make  ||f(x)  —  f(a)  —  £>f(a)(x  —  a)||  as  small  as  we  wish 
by  keeping  ||x  —  a||  appropriately  small.  In  particular, 

||f(x)-f(a)-Df(a)(x-a)||  <  ||x-a|| 

if  ||  x  —  a||  is  sufficiently  small.  To  the  second  term  in  the  right  side  of  inequality 
(10),  we  may  apply  (9),  since  Df(a)  is  an  m  x  n  matrix.  Therefore,  we  see  that  if 
|| x  —  a||  is  made  sufficiently  small, 

||f(x)-f(a)||  <  ||x-a||+Z||x-a||=(l  +  Z)||x-a||.  (11) 

The  constant  K  does  not  depend  on  x.  Thus,  as  x  —>  a,  we  have 

||f(x)  -  f(a)||  -»  0, 

as  desired. 

To  complete  the  proof,  we  establish  inequality  (9).  Writing  out  the  matrix 
multiplication, 


y  =  Bx  = 


b\\%\  +  ^12^2  H  V  bXnxn 

bi\x\  +  b22x2  H  h  b2nx„ 

bm\X\  +  bm2X2  +  ■■■  +  & 


bi  •  x 
b2  -x 


where  b,  denotes  the  ith  row  of  B,  considered  as  a  vector  in  R".  Therefore,  using 
the  Cauchy-Schwarz  inequality, 


||y||  =  ((bi  •  X)2  +  (b2  •  X)2  +  •  •  •  +  (bm  •  X)2) 
<  (||b1||2||x||2  +  ||b2|r||x|r  + 

,2\  1/2 


+  l|bm||2||x||2)1/2 


|b1||2  +  ||b2||2  +  ---  +  ||b„ 


2)' 


Now, 


Consequently, 


=  bjl+bf2  +  ---+bl  =  J2bfJ. 


llbi  ||2  +  ||b2 1|2  +  •  •  •  +  ||b„, ||2  =      lib,  II2  =  E  E bl  =  Rl- 

i=i  i=i  j=\ 

Thus,  ||y||  <  AT||x||,  and  we  have  completed  the  proof  of  Theorem  3.9.  ■ 

Proof  of  Theorem  3.10  First,  we  prove  Theorem  3. 10  for  the  case  where  /  is  a 
scalar-valued  function  of  two  variables.  We  begin  by  writing 

f(x\ ,  x2)  —  f(flu  a2)  =  f(x\ ,  xi)  -  f{a\ ,  x2)  +  f(a\ ,  x2)  -  f(a\,  a2). 

By  the  mean  value  theorem,2  there  exists  a  number  c\  between  a\  and  xi  such 

that 


f{x\,x2)  -  f(ai,x2)  =  fXi(c\,  x2){xx  -  ai) 


2  Recall  that  the  mean  value  theorem  says  that  if  F  is  continuous  on  the  closed  interval  [a ,  b]  and  differen- 
tiable on  the  open  interval  (a,  b),  then  there  is  a  number  c  in  (a ,  b)  such  that  F(b)  —  F(a)  =  F'(c)(b  —  a). 


Chapter  2  |  Differentiation  in  Several  Variables 


and  a  number  c2  between  a2  and  x2  such  that 

f(ai,x2)  -  f(aua2)  =  fX2(au  c2)(x2  -  a2). 

(This  works  because  in  each  case  we  hold  all  the  variables  in  /  constant  except 
one,  so  that  the  mean  value  theorem  applies.)  Hence, 

\f(xux2)  -  f(a\,  a2)  -  fxi{a\,  a2){x\  -  a{)  -  fX2(aua2)(x2  -  a2)\ 
=  \fXl(c\,x2)(x\  -  ai)+  fX2(ai,  c2)(x2  -  a2)  -  fX](au  a2)(x\  -  ax) 
-fX2(au  a2)(x2  -  a2)\ 

<  x2)(.x;i  -  ai)  -  fXl(ax,a2)(xi  -  fli)| 

+  |/x2(ai,  c2)(x2  -  a2)  -  fXl(ai,  a2)(x2  -  a2)\  , 
by  the  triangle  inequality.  Hence, 

\f(xi,x2)  -  f(aua2)  -  fxi(ai,a2)(xi  -  a\)  -  fX2(au  a2)(x2  -  a2)\ 

<  |/*i(Cl,*2)  ~  fxM\'a2)\  \x\  ~  a\\ 

+  \fX2(au  c2)  -  fX2{ax,  a2)\  \x2  -  a2\ 

<  {\fXl(ci,x2)  -  fXl(ai,a2)\  +  \fX2(ai,  c2)  -  fX2(au  a2)\)  ||x  -  a||, 
since,  for  i  =  1,2, 

\Xi  -en\<  ||x  -  a||  =  ((xi  -  ax)2  +  (x2  -  a2)2)1'2. 

Thus, 

|/(xi,  x2)  -  f(au  a2)  -  fxi(aua2)(xi  -  a{)  -  fX2{ax,a2){x2  -  a2)\ 

l|x-a|| 

<  \fxi(c\,x2)- fxl(ai,a2)\  +  \fX2(auc2)-  fX2(ai,a2)\.  (12) 

As  x  — >  a,  we  must  have  that  c,  —>  a\ ,  for  i  =  I,  2,  since  c,  is  between  a,  and  x, . 
Consequently,  by  the  continuity  of  the  partial  derivatives,  both  terms  of  the  right 
side  of  (12)  approach  zero.  Therefore, 

,.     \f(xi,x2)-  f(ax,a2)-  fXl(ai,a2)(xi  -  a{)  -  fX2(a{ ,  a2)(x2  -  a2)\ 

hm  1   =  0 

||x-a|| 

as  desired. 

Exactly  the  same  kind  of  argument  may  be  used  in  the  case  that  /  is  a  scalar- 
valued  function  of  n  variables — the  details  are  only  slightly  more  involved  so 
we  omit  them.  Granting  this,  we  consider  the  case  of  a  vector-valued  function 
f:  R"  — >  Rm.  According  to  Definition  3.8,  we  must  show  that 

||ml|fW-f(a)-Z>fW(x-a)|=0 
||x  -  a|| 

The  component  functions  of  the  expression  appearing  in  the  numerator  may  be 
written  as 

G,  =  Mx)  -  fi(z)  -  D/,(a)(x  -  a),  (14) 

where  fi,  i  =  1, . . . ,  m,  denotes  the  ith  component  function  of  f.  (Note  that,  by 
the  cases  of  Theorem  3.10  already  established,  each  scalar- valued  function  /,  is 


2.3  I  Exercises 


differentiable.)  Now,  we  consider 

||f(x)  -  f(a)  -  Df(a)(x  -  a)||      \\(GU  G2, . . . ,  G„ 


|x-a||  ||x -a|| 

_(G]  +  Gl  +  ---  +  Gjy/2 
Hi -a|| 

<  |Gil  +  |G2|  +  ---  +  |Gm| 
l|x-a|| 

Igil     ]     \G2\     |       |  \G„ 


lix  —  a||      ||x  —  a||  ||x-a|| 

Asx^  a,  eachterm  |G,  |/||x  —  a||  ->  0,  by  definition  of  G,  in  equation  (14)  and 
the  differentiability  of  the  component  functions  f\  of  f .  Hence,  equation  (13)  holds 

and  f  is  differentiable  at  a.  (To  see  that  (G2  H  h  G2  )1/2  <  \G\\-\  hi  Gm  \ , 

note  that 

(IGH  H  h  |Gm|)2  =  |Gi|2  H  h  |G,„|2 

+  2|Gi|  |G2|  +2|Gi|  |G3|  +  ■  ■  ■  +  2|GM_i|  |G„,| 
>  |Gi|2  +  ---  +  |GJ2. 
Then,  taking  square  roots  provides  the  inequality.)  ■ 

Proof  of  Theorem  3.11  In  the  final  paragraph  of  the  proof  of  Theorem  3. 10,  we 
showed  that 

||f(x)-f(a)-Pf(a)(x-a)||  <    |Gt|     |     |G2|     |       |  |G,„| 


||x-a||  llx  —  a||      ||x-a||  l|x-a|| 

where  G,  =  fi(x)  —  /,(a)  —  Z)/}(a)(x  —  a)  as  in  equation  (14).  From  this,  it  fol- 
lows immediately  that  differentiability  of  the  component  functions  f\ ,  . . . ,  fm  at 
a  implies  differentiability  of  fat  a.  Conversely,  for  i  =  1, . . . ,  m, 

||f(x)-f(a)-Df(a)(x-a)||  =  \\(GU  G2,  . . . ,  Gm)||  >  |G,| 


||x-a||  ||x-a||  ||x-a|| 

Hence,  differentiability  of  f  at  a  forces  differentiability  of  each  component 
function.  ■ 


2.3  Exercises 


In  Exercises  1—9,  calculate  df/dx  and  df/dy. 
1-  f(x,y)  =  xy2+x2y 

2.  f(x,  y)  =  e^1 

3.  f(x,  y)  =  sinxy  +  cosjcy 


6.  f(x,  y)  =  \n(x2  +  y2) 

7.  f(x,  y)  =  cosx3y 

'  x  ' 


8.  f(x,  y)  =  In 


4.  f(x,y)  . 


5.  f{x,y)  . 


^  -  y2 
\+x2  +  3/ 

x2-y2 
x2  +  y2 


9.  f(x,  y)  =  xey  +  y  sin(x2  +  y) 

In  Exercises  10—17,  evaluate  the  partial  derivatives  dF/dx, 
dF/dy,  and  dF/dz  for  the  given  functions  F. 

10.  F(x,y,z)  =  x  +  3y-2z 


Chapter  2  |  Differentiation  in  Several  Variables 


x  —  v 

11.  F(x,y,z)=  — ~ 

y  +  z 

12.  F(x,  y,  z)  =  xyz 

13.  F(x,y,z)  =  y/x2  +  y2  +  z2 

14.  F(x,  y,  z)  =  eax  cosfcy  +  eaz  sinbx 

-ic    T7i  ^  x  +  y  +  z 

15.  F(x,  y,  z)  =  — — 

(1  +  x2  +  y2  +  Z2)3/2 

16.  F(x,  y,  z)  =  sinx2y3z4 


17.  F(x,y,z) 


x3  +  yz 


X2  +  Z2  +  1 

FzW  //ze  gradient  V/(a),  where  f  and  a  are  given  in  Exer- 
cises 18-25. 

18.  /(x,  y)  =  x2y  +  ey'x,    a  =  (1,0) 
x  —  y 


19.  f(x,y)  . 


-,    a  =  (2,-1) 


.v2  +  y2  +  V 

20.  /(x,  y,  z)  =  sinxyz,    a  =  (jt,  0,  jr/2) 

21.  /(x,  j,  z)  =  xy  +  y  cosz  —  *  sinyz, 
a  =  (2,-l,jr) 

22.  f(x,  y)  =  c»>  +  In  (x  -  y),     a  =  (2,  1) 

23.  /(x,y,z)=^±2,    a  =  (3, -1,0) 

24.  /(x,y,z)  =  coszln(x  +  y2),    a  =  (e,  0,  jt/4) 

2  _  2 

25.  /(x,y,z)  =    /        Z  ,    ■  =  (-1,2,1) 

yz  +  zz  +  1 

In  Exercises  26-33,  find  the  matrix  Df(  a)  of  partial  derivatives, 
where  f  and  a  are  as  indicated. 

26.  f(x,y)=-,    a  =  (3,  2) 

y 

27.  f(x,y,z)  =  x2+xln(yz),    a  =  (-3,  e,e) 

28.  f(x,  y,  z)  =  (2x  -  3y  +  5z,  x2  +  y,  In  (yz)), 
a  =  (3, -1,-2) 

29.  f(x,  y,  z)  =  (xyz,  ^x2  +  y2  +  z2^), 

a  =  (1,0,  -2) 

30.  f(f)  =  (r,  cos2f,  sin 5/),     a  =  0 

31.  f(x,  y,  z,  id)  =  (3x  -  7y  +  z,  5x  +  2z  -  Sw, 
y-\7z  +  3w),    a  =  (1,2,  3,  4) 

32.  f(x,  y)  =  (x2y,  x  +  y2,  cos  jrxy),     a  =  (2,  -1) 

33.  f(s,  r)  =  (j2,jf,r2),    a  =  (-1,1) 

Explain  why  each  of  the  functions  given  in  Exercises  34—36  is 
differentiable  at  every  point  in  its  domain. 

34.  /(x,  y)  =  xy  —  7x8y2  +  cosx 
x  +  v  +  z 


35.  /(x,y,z): 


36.  f(x,  y)  : 


xy 


xl  +  y4  y 


37.  (a)  Explain  why  the  graph  of  z  =  x3  —  7xy  +  ey  has 

a  tangent  plane  at  (— 1 ,  0,  0). 
(b)  Give  an  equation  for  this  tangent  plane. 

38.  Find  an  equation  for  the  plane  tangent  to  the  graph  of 
z  =  4cosxy  at  the  point  (tt/3,  1,  2). 

39.  Find  an  equation  for  the  plane  tangent  to  the  graph  of 
z  =  ex+y  cosxy  at  the  point  (0,  1,  e). 

40.  Find  equations   for  the  planes  tangent  to  z  = 
x2  —  6x  +  y3    that    are    parallel    to    the  plane 


4x  -  12y  +  j 


7. 


41.  Use  formula  (8)  to  find  an  equation  for  the  hy- 
perplane  tangent  to  the  4-dimensional  paraboloid 
x5  =  10  —  (x\  +  3x|  +  2x|  +  x|)  at  the  point 
(2,-1,  1,3,-8). 

42.  Suppose  that  you  have  the  following  information  con- 
cerning a  differentiable  function  /: 

/(2,  3)  =12,    /(1.98,3)  =  12.1,    f(2,  3.01)  =  12.2. 

( a)  Give  an  approximate  equation  for  the  plane  tangent 
to  the  graph  of  /  at  (2,  3,  12). 

(b)  Use  the  result  of  part  (a)  to  estimate  /(l  .98,  2.98). 

In  Exercises  43—45,  (a)  use  the  linear  function  h(\)  in  Def- 
inition 3.8  to  approximate  the  indicated  value  of  the  given 
function  f.  (b)  How  accurate  is  the  approximation  determined 
in  part  (a)? 

43.  f(x,  y)  =  ex+y,  /(0.1,  -0.1) 

44.  f(x,  y)  =  3  +  cos  itxy,  /(0.98,  0.51) 

45.  f(x,  y,  z)  =  x2  +  xyz  +  y3z,  /(1.01,  1.95,  2.2) 

46.  Calculate  the  partial  derivatives  of 

xi  +  Xj  H  1-  x„ 


/(Xl,X2, 


47.  Let 


fix,  y) 


xy 


,  x„) 


■x2y  +  3x3 


*?  +  xl  +  ■ 


■+x, 


x2  +  y2 
0 


if(x,y)^(0,  0) 
if(x,y)  =  (0,0) 


x2  +  y2  +  z2 


(a)  Calculate  df/dx  and  df/dy  for  (x,  y)/(0,  0). 
(You  may  wish  to  use  a  computer  algebra  system 
for  this  part.) 

(b)  Find  fx(0,  0)  and  /,(0,  0). 

As  mentioned  in  the  text,  if  a  function  F{x)  of  a  single  variable 
is  differentiable  at  a,  then,  as  we  zoom  in  on  the  point  {a,  F(a)), 
the  graph  of  y  =  F(x)  will  "straighten  out"  and  look  like  its 
tangent  line  at  (a,  F(a)).  For  the  differentiable  functions  given 


2.3  I  Exercises  133 


in  Exercises  48-51,  (a)  calculate  the  tangent  line  at  the  indi- 
cated point,  and  (b)  use  a  computer  to  graph  the  function  and 
the  tangent  line  on  the  same  set  of  axes.  Zoom  in  on  the  point 
of  tangency  to  illustrate  how  the  graph  of  y  =  F(x)  looks  like 
its  tangent  line  near  (a,  F(a)). 


^  48.  F(x) 


2x  +  3,     a  =  1 


smx,  a 


^  49.  F{x)  =  x  +  - 

>-3  -  3x2  +  x 


x    —ox    -(-  X 

O  50.  F(x)  =   T—  ,     a  =  0 

xL  +  1 

^  51.  F(x)  =  ln(x2  +  1),     a  =  -1 

52.  (a)  Use  a  computer  to  graph  the  function  F(x)  = 
(x-2)2'\ 

(b)  By  zooming  in  near  x  =  2,  offer  a  geometric  dis- 
cussion concerning  the  differentiability  of  F  at 
x  =2. 

As  discussed  in  the  text,  a  function  fix,  y)  may  have  partial 
derivatives  fx(a,  b)  and  fy(a,  b)  yet  fail  to  be  differentiable  at 
(a,  b).  Geometrically,  if  a  function  fix,  y)  is  differentiable  at 
(a,  b),  then,  aswezoom  in  on  thepoint(a,  b,  f(a,  b)),  thegraph 
of z  =  fix,  y)  will  "flatten  out"  and  look  like  the  plane  given 
by  equation  (4)  in  this  section.  For  the  functions  fix,  y)  given 
in  Exercises  53-57,  (a)  calculate  fx(a,  b)  and  fy(a,  b)  at  the 
indicated  point  (a ,  b)  and  write  the  equation  for  the  plane  given 
by  formula  (4)  of  this  section,  (b)  use  a  computer  to  graph  the 
equation  z  =  f(x,  y)  together  with  the  plane  calculated  in  part 
(a).  Zoom  in  near  the  point  (a,  b,  f(a,  b))  and  discuss  whether 
or  not  f(x,  y)  is  differentiable  at  (a,  b).  (c)  Give  an  analytic 
(i.e.,  nongraphical)  argument  for  your  answer  in  part  (b). 

ft  53.  f(x,  y)  =  x3-xy  +  y2,  (a,  b)  =  (2,  1) 
^>  54.  f{x,  y)  =  Hx  -  \)yf'\     (a,  b)  =  (1,  0) 


^55.  fix,y)  . 
^56.  fix,y)  . 


xy 


(fl,fc)  =  (0,0) 


x2  +  y2  +  1 ' 

fit  3tt  \ 

sinxcosy,     (a,  &)=(—,— ) 
V  6    4  / 


^  57.  fix,y)  =  x2  s'my  +  y2cosx,     (a,A)=^— ,— ) 

58.  Let  gix,  y)  =  ^fxy. 

(a)  Is  g  continuous  at  (0,  0)? 

(b)  Calculate  dg/dx  and  dg/dy  whenxy  ^  0. 

(c)  Show  that  gx(0,  0)  and  g,  (0,  0)  exist  by  supplying 
values  for  them. 

(d)  Are  dg/dx  and  dg/dy  continuous  at  (0,  0)? 

(e)  Does  the  graph  of  z  =  gix,  y)  have  a  tangent  plane 
at  (0,  0)?  You  might  consider  creating  a  graph  of 
this  surface. 

(f)  Is  g  differentiable  at  (0,  0)? 


59.  Suppose  f:  R"  —>  W  is  a  linear  mapping;  that  is, 

f(x)  =  Ax,    where  x  =  (xi,  X2, . . . ,  xn)  e  R" 

and  A  is  an  m  x  n  matrix.  Calculate  Df(x)  and  relate 
your  result  to  the  derivative  of  the  one-variable  linear 
function  fix)  =  ax. 

In  Exercises  60-62  you  will  establish  that  the  matrix  -Df(a)  of 
partial  derivatives  of  the  component  functions  ofiis  uniquely 
determined  by  the  limit  equation  in  Definition  3.8. 

60.  Let  X  be  an  open  set  in  R",  let  aeX,  and  let  F:ic 
R"  -±  R".  Show  that 


lim  ||F(x)||  =  0 


lim  F(x)  =  0. 


61.  Let  X  be  an  open  set  in  R",  let  ael,  and  let  f:XC 
R"  — >•  R" .  Suppose  that  A  and  B  are  m  x  n  matrices 
such  that 


lim 

x— >a 


lim 


||f(x)-[f(a)  +  A(x-a)]| 

llx  —  a|| 
f(x)  -  [f(a)  +  5(x  -  a)]  || 


x^a  ||  x  —  a|| 

(a)  Use  Exercise  60  to  show  that 

hmCS-AKx-a, 
x->-a       Hx  —  all 


0. 


(b)  Write  x  —  a  as  fh,  where  h  is  a  nonzero  vector  in 
R" .  First  argue  that 


lim 

x-»a 


(S  -  A)(x  -  a) 


0  implies 


lim(B-A)W 
t-o  \\th\\ 

and  then  use  this  result  to  conclude  that  A  =  B. 
(Hint:  Break  into  cases  where  t  >  0  and  where 
t  <  0.) 

62.  Let  X  be  an  open  set  in  R",  let  aeX,  and  let  f:XC 
R"  — >•  R" .  Suppose  that  A  is  an  m  x  n  matrix  such 
that 


lim 

x->a 


||f(x)-[f(a)  +  A(x-a)] 


0. 


In  this  problem  you  will  establish  that  A  =  Df(a). 
(a)  Define  F:ICR"->  R"  by 


F(x)  = 


f(x)  -  f(a)  -  A(x  -  a) 


Identify  the  ;th  component  function  us- 
ing component  functions  of  f  and  parts  of  the 
matrix  A. 

(b)  Note  that  under  the  assumptions  of  this  problem 
and  Exercise  60,  we  have  that  limx^a  F(x)  =  0. 


1  34       Chapter  2  |  Differentiation  in  Several  Variables 


First  argue  that,  for  i=l,...,m,  we  have 
limx^a  Fi(x)  =  0.  Next,  argue  that 

lim  F,(x)  =  0    implies    lim  F,-(a  +  he:)  =  0, 

where  denotes  the  standard  basis  vector 
(0, 1,  ...,0)forR". 


(c)  Use  parts  (a)  and  (b)  to  show  that  atj 


M. 


(a), 


where  a,;  denotes  the  ijth  entry  of  A.  (Hint:  Break 
into  cases  where  h  >  0  and  where  h  <  0.) 


2.4   Properties;  Higher-order  Partial 
Derivatives 

Properties  of  the  Derivative   

From  our  work  in  the  previous  section,  we  know  that  the  derivative  of  a  function 
f :  X  c  R"  R'"  can  be  identified  with  its  matrix  of  partial  derivatives.  We  next 
note  several  properties  that  the  derivative  must  satisfy.  The  proofs  of  these  results 
involve  Definition  3.8  of  the  derivative,  properties  of  ordinary  differentiation,  and 
matrix  algebra. 


PROPOSITION  4.1  (Linearity  of  differentiation)  Let  f,g:Jfc  R"  -»  Rffl 
be  two  functions  that  are  both  differentiable  at  a  point  aeX,  and  let  c  e  R  be 
any  scalar.  Then 

1.  The  function  h  =  f  +  g  is  also  differentiable  at  a,  and  we  have 

Z)h(a)  =  D(f  +  g)(a)  =  Df(a)  +  £»g(a). 

2.  The  function  k  =  cf  is  differentiable  at  a  and 

£>k(a)  =  £>(cf)(a)  =  c£>f(a). 


EXAMPLE  1  Let  f  and  g  be  defined  by  f(x,  y)  =  (x  +  y,  xy  siny,  y/x)  and 
g(x,  y)  =  (x2  +  y2,  yexy ,  2x3  —  7y5).  We  have 


Df(x,  y)  = 


1 

y  siny 

L  -3V*2 


and 


Dg(x,y)  = 


2x 

y2exy 
6x2 


1 

x  siny  +  xy  cosy 
l/x 


2y 

ex>'  +  xye^^ 
-35y4 


Thus,  by  Theorem  3.10,  f  is  differentiable  on  R2  —  {y-axis}  and  g  is  differentiable 
on  all  of  R2.  If  we  let  h  =  f  +  g,  then  part  1  of  Proposition  4. 1  tells  us  that  h  must 
be  differentiable  on  all  of  its  domain,  and 

Dh(x,y)  =  Df(x,y)  +  Dg(x,y) 

2x  +  \  2y  +  1 

y  sin  y  +  y2exy      x  sin  y  +  xy  cos  y  +  exy  +  xyexy 
6x2  -y/x2  l/x-  35y4 


2.4  |  Properties;  Higher-order  Partial  Derivatives 


Note  also  that  the  function  k  =  3g  must  be  differentiable  everywhere  by  part  2 
of  Proposition  4.1.  We  can  readily  check  that  Dk(x,  y)  =  3Dg(x,  y):  We  have 

k(x,  y)  =  (3x2  +  3y2,  3yexy ,  6x3  -  21y5). 

Hence, 


£>k(x,  y)  = 


6.v 

6y 

3exy  +  3xye 

18x2 

-105y4 

2x 

2y 

y2exy 

exy  +  xyexy 

6x2 

-35y4 

=  3 


=  3£>g(x,  y).  4 

Due  to  the  nature  of  matrix  multiplication,  general  versions  of  the  product 
and  quotient  rules  do  not  exist  in  any  particularly  simple  form.  However,  for 
scalar-valued  functions,  it  is  possible  to  prove  the  following: 

PROPOSITION  4.2  Let  /,g:lcR"^Rbe  differentiable  ataeX  Then 

1.  The  product  function  fg  is  also  differentiable  at  a,  and 

D(fgM  =  s(a)D/(a)  +  /(a)Z)g(a). 

2.  If  g(a)  /  0,  then  the  quotient  function  f/g  is  differentiable  at  a,  and 

g(a)D/(a)  -  /(a)Z)g(a) 


D(//g)(a)  = 


g(a)2 


(/#)(*>  y,  z)  =  (xyz  +  2yz2  -  xz2)exy, 


EXAMPLE  2  If  f(x,  y,  z)  =  z<?x-v  and  g(x,  y,  z)  =  xy  +  2yz  -  xz,  then 
so  that 

D(fg)(x,y,z) 


(yz  -  z2)exy  +  (xyz  +  2yz2  -  xz2)yexy 
(xz  +  2z2)exy  +  (xyz  +  2yz2  -  xz2)xexy 
(xy  +  4yz  —  2xz)ex>' 


-i  t 


Also,  we  have 
and 


Df(x,y,z)=  [yzexy   xZexy  exy] 
Dg(x,y,z)=[y  -  z  x  +  2z  2y-x], 


so  that 


g(x,  y,  z)Df(x,  y,  z)  +  f(x,  y,  z)Dg(x,  y,  z) 

(xy2z  +  2y2z2  —  xyz2)exy 
(x2yz  +  2xyz2  -  x2z2)exy 
(xy  +  2yz  -  xz)exy 


(yz  -  z2)exy 
(xz  +  2z2)exv 
(2yz  -  xz)exy 


-1  T 


Chapter  2  |  Differentiation  in  Several  Variables 


=  e 


xy2z 
x2yz  - 


-  2y2z2  -  xyz2  +  yz-  z2 
2xyz2  -  x2z2  +  xz  +  2z2 
xy  +  4yz  —  2xz 


which  checks  with  part  1  of  Proposition  4.2.  (Note:  The  matrix  transpose  is  used 
simply  to  conserve  space  on  the  page.)  ♦ 


The  product  rule  in  part  1  of  Proposition  4.2  is  not  the  most  general 
result  possible.  Indeed,  if  /:ICR"->R  is  a  scalar- valued  function  and 
g:lcR"^  R'"  is  a  vector- valued  function,  then  if  /  and  g  are  both  differ- 
entiate at  a  e  X,  so  is  fg,  and  the  following  formula  holds  (where  we  view  g(a) 
as  an  m  x  1  matrix): 

£>(/g)(a)  =  g(a)D/(a)  +  /(a)Dg(a). 


Partial  Derivatives  of  Higher  Order  

Thus  far  in  our  study  of  differentiation,  we  have  been  concerned  only  with  partial 
derivatives  of  first  order.  Nonetheless,  it  is  easy  to  imagine  computing  second- 
and  third-order  partials  by  iterating  the  process  of  differentiating  with  respect  to 
one  variable,  while  all  others  are  held  constant. 

EXAMPLE  3  Let  f(x,  y,  z)  =  x2y  +  y2z.  Then  the  first-order  partial  deriva- 
tives are 

9/  „  9/  2  „  9/2 
—  =  2xy,     —  =  x  +  2yz,     and    —  =  y  , 

ax  ay  az 

The  second-order  partial  derivative  with  respect  to  x,  denoted  by  d2f/dx2  or 
fxx(x,  y,  z),  is 

d2f      9  (df\      3  .„    ,  „ 

— —  =  —  —  =  — (2xy)  =  2y- 

dx2      dx  \dx )  dx 
Similarly,  the  second-order  partials  with  respect  to  y  and  z  are,  respectively, 

92/      9  /3A  d 


dy2      dy  \dy J  dy^ 


and 

d2f      3  (df 


dz2       dz  \  dz  J      dz  ^ 

There  are  more  second-order  partials,  however.  The  mixed  partial  derivative 
with  respect  to  first  x  and  then  y,  denoted  d2  f /dydx  or  fxy(x,  y,  z),  is 

92/       9  /3/\      3  , 

=  — (2xy)  =  2x. 


dydx      dy  \dx )  dy 

There  are  five  more  mixed  partials  for  this  particular  function:  d2f/dxdy, 
d2f/dzdx,  d2  f /dxdz,  d2  f /dzdy,  and  d2  f /dydz.  Compute  each  of  them  to  get 
a  feeling  for  the  process.  ♦ 

In  general,  if  /:  X  C  R"  R  is  a  (scalar- valued)  function  of  n  variables,  the 
&th-order  partial  derivative  with  respect  to  the  variables  (in  that 


2.4  |  Properties;  Higher-order  Partial  Derivatives 


order),  where  ii ,  1*2,  -..,£*  are  integers  in  the  set  {1,  2, ....  n}  (possibly  repeated), 
is  the  iterated  derivative 

dxh- ■  ■  dxhdxil      dxk  dxkdxil 

Equivalent  (and  frequently  more  manageable)  notation  for  this  /rth-order  partial 
is 

Note  that  the  order  in  which  we  write  the  variables  with  respect  to  which  we 
differentiate  is  different  in  the  two  notations:  In  the  subscript  notation,  we  write 
the  differentiation  variables  from  left  to  right  in  the  order  we  differentiate,  while 
in  the  3 -notation,  we  write  those  variables  in  the  opposite  order  (i.e.,  from  right 
to  left). 

EXAMPLE  4   Let  f(x,  y,  z,  w)  =  xyz  +  xy2w  —  cos(x  +  zw).  We  then  have 

d2f        9    9  , 
fyw(x,y,z,w)=  — —  =  — — (xyz  +  xy  w  -  cos(x  +  zw)) 
away      aw  ay 

9 

=  — (xz  +  2xyw)  =  2xy, 
dw 


and 


d2f        9    9  2 
fwy(x,  y,  z,  w)  =  ——  =  —  —(xyz  +  xy  w  -  cos(x  +  zw)) 
ayaw      ay  aw 


9 

9^ 


(xy2  +  z  sin(x  +  zw))  =  2xy. 


♦ 


Although  it  is  generally  ill-advised  to  formulate  conjectures  based  on  a  single 
piece  of  evidence,  Example  4  suggests  that  there  might  be  an  outrageously  simple 
relationship  among  the  mixed  second  partials.  Indeed,  such  is  the  case,  as  the  next 
result,  due  to  the  1 8th-century  French  mathematician  Alexis  Clairaut,  indicates. 


THEOREM  4.3  Suppose  that  X  is  open  in  R"  and  /:XcR"^R  has  con- 
tinuous first-  and  second-order  partial  derivatives.  Then  the  order  in  which  we 
evaluate  the  mixed  second-order  partials  is  immaterial;  that  is,  if  i\  and  12  are  any 
two  integers  between  1  and  n,  then 

J2J_  =  J2J_ 
dxj1  dxj2       dxi2  dxj1 


A  proof  of  Theorem  4.3  is  provided  in  the  addendum  to  this  section.  We 
also  suggest  a  second  proof  (using  integrals!)  in  Exercise  4  of  the  Miscellaneous 
Exercises  for  Chapter  5. 

It  is  natural  to  speculate  about  the  possibility  of  an  analogue  to  Theorem  4.3 
for  &th-order  mixed  partials.  Before  we  state  what  should  be  an  easily  anticipated 
result,  we  need  some  terminology. 


Chapter  2  |  Differentiation  in  Several  Variables 


DEFINITION  4.4  Assume  X  is  open  in  R".  A  scalar-valued  function 
/:XcR"^R  whose  partial  derivatives  up  to  (and  including)  order  at 
least  k  exist  and  are  continuous  on  X  is  said  to  be  of  class  Ck.  If  /  has 
continuous  partial  derivatives  of  all  orders  on  X,  then  /  is  said  to  be  of  class 
C°°,  or  smooth.  A  vector-valued  function  f:  X  c  R"  ->  R'"  is  of  class  Ck 
(respectively,  of  class  C°°)  if  and  only  if  each  of  its  component  functions  is 
of  class  Ck  (respectively,  C°°). 


THEOREM  4.5  Let  /:  X  c  R"  ->  R  be  a  scalar-valued  function  of  class  Ck. 
Then  the  order  in  which  we  calculate  any  £th-order  partial  derivative  does  not 
matter:  If  . . . ,  /<-)  are  any  k  integers  (not  necessarily  distinct)  between  1  and 
n,  and  if  (ji,  ... ,  jk)  is  any  permutation  (rearrangement)  of  these  integers,  then 

dkf       =  dkf 
dxj{  ■  ■  ■  dxjt       dxjj  ■  ■  ■  dxjt 


EXAMPLE  5  If  f(x,  y,  z,  w)  =  x2weyz  —  zexw  +  xyzw,  then  you  can  check 
that 


95f  d5f 
2eyz(yz  +  \) 


dxdwdzdydx  dzdydwd2x' 
verifying  Theorem  4.5  in  this  case. 


Addendum:  Two  Technical  Proofs   

Proof  of  Part  1  of  Proposition  4.1 

Step  1.  We  show  that  the  matrix  of  partial  derivatives  of  h  is  the  sum  of 
those  of  f  and  g.  If  we  write  h(x)  as  (hi(x),  /i2(x),  . . . ,  hm(x))  (i.e.,  in  terms  of 
its  component  functions),  then  the  ijth  entry  of  £>h(a)  is  dhj/dxj  evaluated  at  a. 
But  hj(x)  =  fi(x)  +  gi(x)  by  definition  of  h.  Hence, 

dh;        9  dfj  dg; 

— ^  =  —(/;■«  +  oo)  =  -f-  + 

dXj      dXj  axj  oxj 

by  properties  of  ordinary  differentiation  (since  all  variables  except  xj  are  held 
constant).  Thus, 

^—  (a)=  —  (a)+  —  (a), 

axj  axj  axj 


and  therefore, 


Z)h(a)  =  £>f(a)  +  Dg(a). 


Step  2.  Now  that  we  know  the  desired  matrix  of  partials  exists,  we  must 
show  that  h  really  is  differentiable;  that  is,  we  must  establish  that 

Hm  ||h(x)-[h(a)  +  Ph(a)(x-a)]||  =  Q 


2.4  |  Properties;  Higher-order  Partial  Derivatives 


(a,  b  +  Ay) 

(a  +  Ax,  b  +  Ay) 

• 

• 

+ 

+ 

• 

(a'b) 

(a  +  Ax,  b) 

Figure  2.55  To  construct  the 
difference  function  D  used  in  the 
proof  of  Theorem  4.3,  evaluate  / 
at  the  four  points  shown  with  the 
signs  as  indicated. 


As  preliminary  background,  we  note  that 

||h(x)-[h(a)  +  /)h(a)(x-a)]|| 
l|x-a|| 

||f(x)  +  g(x)  -  [f(a)  +  g(a)  +  £>f(a)(x  -  a)  +  Dg(a)(x  -  a)]|| 

II* -a|| 

||(f(x)  -  [f(a)  +  Df(a)(x  -  a)])  +  (g(x)  -  [g(a)  +  Dg(a)(x  -  a)])|| 

II* -a|| 

<  ||f(x)-[f(a)  +  Df(a)(x-a)]||  |  ||g(x)  -  [g(a)  +  Dg(a)(x  -  a)]|| 


x-a 


x  -  a 


by  the  triangle  inequality,  formula  (2)  of  §1.6.  To  show  that  the  desired  limit 
equation  for  h  follows  from  the  definition  of  the  limit,  we  must  show  that  given 
any  e  >  0,  we  can  find  a  number  S  >  0  such  that 


.,n     ,        ,     .  t,      ||h(x)-[h(a)  +  £>h(a)(x-a)]|| 

if  0  <  || x  -  a||  <  8,  then   <  e. 


x  -  a 


(1) 


Since  f  is  given  to  be  differentiable  at  a,  this  means  that  given  any  e\  >  0,  we  can 
find  S\  >  0  such  that 


•fn     ,i        ,i     *   a     llf(x)-[f(a)  +  flf(a)(x-a)]| 
if  0  <  || x  —  a||  <  Si,  then   <  ei. 


x-a 


(2) 


Similarly,  differentiability  of  g  means  that  given  any  62  >  0,  we  can  find  a  82  >  0 
such  that 


Tn     „        „     .   t,      ||g(x)-[g(a)  +  JDg(a)(x-a)]| 
if  0  <  || x  —  a||  <  h,  then   <  €2. 


x  -  a 


(3) 


Now  we're  ready  to  establish  statement  (1).  Suppose  e  >  0  is  given.  Let  81 
and  82  be  such  that  (2)  and  (3)  hold  with  ei  =  €2  =  e/2.  Take  8  to  be  the  smaller 
of  81  and  82-  Hence,  if  0  <  ||x  —  a||  <  <5,  then  both  statements  (2)  and  (3)  hold 
(with  €1  =  €2  =  e/2)  and,  moreover, 

||h(x)-[h(a)  +  Ph(a)(x-a)]||      ||f(x)  -  [f(a)  +  Df(a)(x  -  a)]|| 


x-a 


Hi -a|| 

|g(x)-[g(a)  +  Z)g(a)(x-a)]| 


+ 


x-a 


<el+e2 
e  e 
=  2  +  2=€- 

That  is,  statement  (1)  holds,  as  desired.  ■ 

Proof  of  Theorem  4.3  For  simplicity  of  notation  only,  we'll  assume  that  /  is  a 
function  of  just  two  variables  (x  and  y).  Let  the  point  (a ,  b)  e  R2  be  in  the  interior 
of  some  rectangle  on  which  fx,  fy,  fxx,  fyy,  fxy,  and  fyx  are  all  continuous. 
Consider  the  following  "difference  function."  (See  Figure  2.55.) 

D(Ax,  Ay)  =  f(a  +  Ax,  b  +  Ay)  -  f(a  +  Ax,  b) 
-f(a,b+  Ay)  +  f(a,b). 


1  40       Chapter  2  |  Differentiation  in  Several  Variables 


(a,  b  +  Ay)     (a  +  Ax,  b  +  Ay) 


R 

{'d 

X 


(a,b) 


(a  +  Ax,  b) 


Figure  2.56  Applying  the  mean 
value  theorem  twice. 


Our  proof  depends  upon  viewing  this  function  in  two  ways.  We  first  regard  D  as 
a  difference  of  vertical  differences  in  /: 

D(Ax,  Ay)  =  [f(a  +  Ax,b  +  Ay)  -  f{a  +  Ax,  b)] 

-  [f(a,  b  +  Ay)  -  f  (a,  b)] 

=  F(a  +  Ax)  -  F(a). 

Here  we  define  the  one-variable  function  F(x)  to  be  f(x,  b  +  Ay)  —  f(x,  b).  As 
we  will  see,  the  mixed  second  partial  of  /  can  be  found  from  two  applications  of 
the  mean  value  theorem  of  one-variable  calculus.  Since  /  has  continuous  partials, 
it  is  differentiable.  (See  Theorem  3. 10.)  Hence,  F  is  continuous  and  differentiable, 
and,  thus,  the  mean  value  theorem  implies  that  there  is  some  number  c  between 
a  and  a  +  Ax  such  that 


D(Ax,  Ay)  =  F(a  +  Ax)  -  F(a)  =  F'(c)Ax. 


(4) 


Now  F'{c)  =  fx{c,  b  +  Ay)  —  fx(c,  b).  We  again  apply  the  mean  value  theorem, 
this  time  to  the  function  fx(c,  y).  (Here,  we  think  of  c  as  constant  and  y  as  the 
variable.)  By  hypothesis  fx  is  differentiable  since  its  partial  derivatives,  fxx  and 
fxy,  are  assumed  to  be  continuous.  Consequently,  the  mean  value  theorem  applies 
to  give  us  a  number  d  between  b  and  b  +  Ay  such  that 


F'(c)  =  fx(c,  b  +  A.y)  -  fx(c,  b)  =  fxy(c,  d)Ay. 
Using  equation  (5)  in  equation  (4),  we  have 

D(Ax,  Ay)  =  F'(c)Ax  =  fxy(c,  d)AyAx. 


(5) 


The  point  (c,  d)  lies  somewhere  in  the  interior  of  the  rectangle  R  with  vertices 
(a,  b),  {a  +  Ax,  b),  (a,  b  +  Ay),  (a  +  Ax,  b  +  Ay),  as  shown  in  Figure  2.56. 
Thus,  as  (Ax,  Ay)  ->  (0,  0),  we  have  (c,  d)  — >  (a,  b).  Hence,  it  follows  that 


fxy(c,  d)  ->  fxy(a,  b)    as    (Ax,  Ay) 
since  fxy  is  assumed  to  be  continuous.  Therefore, 


(0,  0), 


D(Ax,  Ay) 

fxJa,b)=       lim       fxJc,d)=       lim   . 

(Ajc,A)0-»-(o,O)    "  (Ax,Ajo^(0,0)  AyAx 

On  the  other  hand,  we  could  just  as  well  have  written  D  as  a  difference  of 
horizontal  differences  in  /: 

£>(Ax,  Ay)  =  [/(a  +  Ax,  b  +  Ay)  -  f(a,  b  +  Ay)] 

-  [f(a  +  Ax,  b)  -  f(a,  b)] 

=  G(b  +  Ay)  -  G(b). 

Here  G(y)  =  f(a  +  Ax,  y)  —  f(a,  y).  As  before,  we  can  apply  the  mean  value 
theorem  twice  to  find  that  there  must  be  another  point  (c,  d)  in  R  such  that 


D(Ax,  Ay)  =  G'(d)Ay  =  fyx(c,  d)AxAy. 


Therefore, 

fyx(a,  b)  = 


lim       fvx(c,  d) 

(Ax,Av)^(0,0)  " 


lim 

(Ajt,Ay)->-(0,0) 


D(Ax,  Ay) 
Ax  Ay 


Because  this  is  the  same  limit  as  that  for  fxy(a,  b)  just  given,  we  have  established 
the  desired  result.  ■ 


2.4  I  Exercises  141 


2.4  Exercises 


In  Exercises  1—4,  verify  the  sum  rule  for  derivative  matrices 
(i.e.,  part  1  of  Proposition  4.1)  for  each  of  the  given  pairs  of 
functions: 

1.  f(x,  y)  —  xy  +  cosx,    g(x,  y)  =  sin(xy)  +  y3 

2.  f{x,  y)  =  (ex+y,xey),     g(x,  y)  =  Qn(xy),  yex) 

3.  f(x,  y,  z)  =  (x  sin  y  +  z,  yez  —  3x2),     g(x,  y,  z)  = 
(x3  cosx,  xyz) 

4.  f(x,  y,  z)  =  (xyz2,  xe~y ,  y  sinxz),     g(x,  y,  z)  = 
(x  —  y,  x2  +  y2  +  z2,  ln(xz  +  2)) 

Verify  the  product  and  quotient  rules  (Proposition  4.2)  for  the 
pairs  of  functions  given  in  Exercises  5—8. 

x 

5-  f(x,  y)  =  x2y  +  y3,     g(x,  y)  =  - 

y 

6.  f(x,y)  =  exy,     g(x,  y)  =  x  sin2y 

7-  fix,  y)  =  3xy  +  y5 ,     g(x,  y)  =  x3  -  2xy2 

8.  f(x,y,z)  =  xcos(yz), 

g(x,  y,  z)  =  x2  +  x9y2  +  y2z3  +  2 

For  the  functions  given  in  Exercises  9-1 7  determine  all  second- 
order  partial  derivatives  (including  mixed partials). 


9. 

fix,  y)  = 

x3y7  +  3xy2  —  Ixy 

10. 

fix,  y)  = 

cos  (xy) 

11. 

fix,  y)  = 

ey/x  _  ye-x 

12. 

fix,  y)  = 

siny7*2  +  y2 

13. 

fix,  y)  = 

1 

sin2  x  +  2ey 

14. 

fix,y)  = 

15. 

fix,  y)  = 

y  sin  x  —  x  cos  y 

16. 

fix,  y)  = 

17. 

fix,y)  = 

x2ey  +  e2z 

18. 

fix,  y,  z) 

x  —  y 
y  +  z 

19. 

f(x,y,z) 

=  x2yz  +  xy2z  +  xyz2 

20. 

fix,  y,  z) 

=  exyz 

21. 

fix,  y,  z) 

=  eax  sin  y  +  ehx  cos  z 

22. 

Consider 

the  function  F(x,y,z) 

y3Z5  —  Ixyz. 


(a)  Find  Fxx,  Fyy,  and  F-z. 

(b)  Calculate  the  mixed  second-order  partials  Fxy, 
Fyx,  Fxz,  F^x,  Fyz,  and  Fzy,  and  verify  Theorem 
4.3. 


(c)  Is  Fxyx  =  Fxxyl  Could  you  have  known  this  with- 
out resorting  to  calculation? 

(d)  Is  Fxyz  =  Fyzx7 

23.  Let   f(x,  y)  =  ye3x.   Give   general   formulas  for 

d"f/dxn  and  dnf/dyn,  where  n  >  2. 

24.  Let  f(x,y,z)  =  xe2y  +  yeiz+ze~x.  Give  general 
formulas  for  d"f/dx",  dnf/dyn,  and  d"f/dz",  where 
n  >  1. 

25.  Let  f(x,  y,  z)  =  In  ^— ^.  Give  general  formulas  for 

d"f/dx",  d"f/dyn,  and  dnf/dz",  where  n  >  1.  What 
can  you  say  about  the  mixed  partial  derivatives? 

26.  Let  f(x,  y,  z)  =  x1y2z3  -  2x4yz. 

(a)  What  is  d4f/dx2dydzl 

(b)  What  is  dsf/dx*dydzl 

(c)  What  is  915//9*133v9z? 

27.  Recall  from  §2.2  that  a  polynomial  in  two  variables  x 
and  y  is  an  expression  of  the  form 

d 

p(x,  y)=Yl  ckixky', 

k.l=0 

where  c«  can  be  any  real  number  for  0  <  k,  I  <  d.  The 
degree  of  the  term  cuxky[  when  cu  ^  0  is  k  +  I  and 
the  degree  of  the  polynomial  p  is  the  largest  degree 
of  any  nonzero  term  of  the  polynomial  (i.e.,  the  largest 
degree  of  any  term  for  which  cu  /  0).  For  example, 
the  polynomial 

p(x,  y)  =  7*y  +  2x2y3  -  3x4  -  5xy3  +  1 

has  five  terms  of  degrees  15,  5,  4,  4,  and  0.  The  de- 
gree of  p  is  therefore  1 5 .  (Note:  The  degree  of  the  zero 
polynomial  p(x,  y)  =  0  is  undefined.) 

(a)  If  p(x,  y)  =  8jc7v10  —  9x2y  +  2x,  what  is  the  de- 
gree of  dp/dx7  dp/dyi  d2p/dx27  32p/dy2? 
d2p/dxdy? 

(b)  If  p(x,  y)  =  Sx2y  +  2x3y,  what  is  the  degree  of 
dp/dx?  dp/dyl  d2p/dx27  d2p/dy2?  d2p/dxdy? 

(c)  Try  to  formulate  and  prove  a  conjecture  relating 
the  degree  of  a  polynomial  p  to  the  degree  of  its 
partial  derivatives. 

28.  The  partial  differential  equation 

32/     3*7  32/ 

— -  H          H          =  0 

dx2      dy2  dz2 

is  known  as  Laplace's  equation,  after  Pierre  Simon 
de  Laplace  (1749-1827).  Any  function  /  of  class  C2 


1  42       Chapter  2  |  Differentiation  in  Several  Variables 


that  satisfies  Laplace's  equation  is  called  a  harmonic 
function.3 

(a)  Is  f(x,  y,  z)  =  x2  +  y2  —  2z2  harmonic?  What 
about  f(x,  y,  z)  =  x2  -  y2  +  z21 

(b)  We  may  generalize  Laplace's  equation  to  functions 
of  n  variables  as 

32/      32/  92/  „ 

— y  +  — T +  ■■■+— V=0. 

ox{      9xj  9jc„ 

Give  an  example  of  a  harmonic  function  of  n  vari- 
ables, and  verify  that  your  example  is  correct. 

29.  The  three-dimensional  heat  equation  is  the  partial  dif- 
ferential equation 

/d2T      d2T      d2T\  _  dT 
\~d~x2  +  ~d~y2  +  Ik2)  ~  ~dt' 

where  k  is  a  positive  constant.  It  models  the  tempera- 
ture T(x,  y,  z,  t)  at  the  point  (x,  y,  z)  and  time  t  of  a 
body  in  space. 

(a)  We  examine  a  simplified  version  of  the  heat  equa- 
tion. Consider  a  straight  wire  "coordinatized"  by 
x.  Then  the  temperature  T(x,  t)  at  time  t  and  po- 
sition x  along  the  wire  is  modeled  by  the  one- 
dimensional  heat  equation 

d2T  _  dT 
Jx2  ~  ~dt' 

Show  that  the  function  T(x,  t)  =  e~kt  cosjc  satis- 
fies this  equation.  Note  that  if  t  is  held  constant  at 
value  ?o,  then  T(x,  to)  shows  how  the  temperature 
varies  along  the  wire  at  time  to.  Graph  the  curves 
z  =  T(x,  to)  for  to  =  0,1,  10,  and  use  them  to  un- 
derstand the  graph  of  the  surface  z  =  T(x,t)  for 
t  >  0.  Explain  what  happens  to  the  temperature  of 
the  wire  after  a  long  period  of  time. 

(b)  Show  that  T(x,  y,  t)  =  e~k'(cosx  +  cos y)  satis- 
fies the  two-dimensional  heat  equation 

id2T      d2T\  _  dT 
[jx*  +  Jy2)  =  It' 


Graph  the  surfaces  given  by  z  =  T(x,  y,  to),  where 
to  =  0,  1,  10.  If  we  view  the  function  T(x,  y,  t)  as 
modeling  the  temperature  at  points  (x,  y)  of  a  flat 
plate  at  time  t,  then  describe  what  happens  to  the 
temperature  of  the  plate  after  a  long  period  of  time. 

(c)  Now  show  that  T(x,  y,  z,  t)  =  e~kt(cosx  + 
cos  y  +  cos  z)  satisfies  the  three-dimensional  heat 
equation. 

30.  Let 

[-^(4x4)  30^(0,0) 
/(je>y)=J      \x2  +  y2J 

[  0  if(x,y)  =  (0,0) 

(a)  Find  fx (x,y) and  /,(* ,  y) for (x,  y)  ?  (0,  0).  (You 
will  find  a  computer  algebra  system  helpful.) 

(b)  Either  by  hand  (using  limits)  or  by  means  of 
part  (a),  find  the  partial  derivatives  /r(0,  y)  and 
/,(x,0). 

(c)  Find  the  values  of  fxy(0,  0)  and  /y.,-(0,  0).  Recon- 
cile your  answer  with  Theorem  4.3. 

A  surface  that  has  the  least  surface  area  among  all  surfaces 
with  a  given  boundary  is  called  a  minimal  surface.  Soap  bub- 
bles are  naturally  occurring  examples  of  minimal  surfaces.  It 
is  a  fact  that  minimal  surfaces  having  equations  of  the  form 
z  =  fix,  y)  (where  f  is  of  class  C2)  satisfy  the  partial  differ- 
ential equation 

(1  +  l])  ZXX  +  (1  +  Zl)  Zyy  =  2ZxZyZxy.  (6) 

Exercises  31—33  concern  minimal  surfaces  and  equation  (6). 

31 .  Show  that  a  plane  is  a  minimal  surface. 

32.  Scherk's  surface  is  given  by  the  equation  ez  cos  y  = 
cosx. 

(a)  Use  a  computer  to  graph  a  portion  of  this  surface. 

(b)  Verify  that  Scherk's  surface  is  a  minimal  surface. 

33.  One  way  to  describe  the  surface  known  as  the  helicoid 
is  by  the  equation  x  =  y  tan  z. 

(a)  Use  a  computer  to  graph  a  portion  of  this  surface. 

(b)  Verify  that  the  helicoid  is  a  minimal  surface. 


2.5   The  Chain  Rule 

Among  the  various  properties  that  the  derivative  satisfies,  one  that  stands  alone 
in  both  its  usefulness  and  its  subtlety  is  the  derivative's  behavior  with  respect 
to  composition  of  functions.  This  behavior  is  described  by  a  formula  known  as 


3  Laplace  did  fundamental  and  far-reaching  work  in  both  mathematical  physics  and  probability  theory. 
Laplace's  equation  and  harmonic  functions  are  part  of  the  field  of  potential  theory,  a  subject  that  Laplace 
can  be  credited  as  having  developed.  Potential  theory  has  applications  to  such  areas  as  gravitation,  elec- 
tricity and  magnetism,  and  fluid  mechanics,  to  name  a  few. 


2.5  |  The  Chain  Rule  143 


the  chain  rule.  In  this  section,  we  review  the  chain  rule  of  one- variable  calculus 
and  see  how  it  generalizes  to  the  cases  of  scalar-  and  vector-valued  functions  of 
several  variables. 

The  Chain  Rule  for  Functions  of  One  Variable:  A  Review 

We  begin  with  a  typical  example  of  the  use  of  the  chain  rule  from  single-variable 
calculus. 

EXAMPLE  1  Let  f(x)  =  sinx  and  x(t)  =  t3  +  t.  We  may  then  construct  the 
composite  function  /(x(r))  =  sin(f3  +  t).  The  chain  rule  tells  us  how  to  find  the 
derivative  of/oi  with  respect  to  t : 

(/  o  jc)'(r)  =  -^-(sin(73  +  0)  =  (cos(?3  +  t))(3t2  +  1). 
at 

Since  x  =  r3  +  t,  we  have 

(/  o  x)'(t)  =  isinjc)  ■  ^{t1  +  0  =  f{x)  ■  x'(t).  ♦ 
ax  at 

In  general,  suppose  X  and  T  are  open  subsets  of  R  and  /:XcR^  R 
and  x :  T  C  R  — >•  R  are  functions  defined  so  that  the  composite  function 
/  o  x :  T  ->  R  makes  sense.  (See  Figure  2.57.)  In  particular,  this  means  that  the 
range  of  the  function  x  must  be  contained  in  X,  the  domain  of  /.  The  key  result 
is  the  following: 


/ 


-i  )  ( — h 


X 

R  R  R 

Figure  2.57  The  range  of  the  function  x  must  be  contained  in  the  domain  X  of  /  in 
order  for  the  composite  /  o  x  to  be  defined. 


THEOREM  5.1  (The  chain  rule  in  one  variable)  Under  the  preceding  as- 
sumptions, if  x  is  differentiable  at  to  6  Tand  /isdifferentiableat.To  =  x(to)  6  X, 
then  the  composite  /  o  x  is  differentiable  at  to  and  moreover, 

(/  o  x)'(to)  =  f\xo)x'(to).  (1) 

A  more  common  way  to  write  the  chain  rule  formula  in  Theorem  5.1  is 
df  df  dx 

Mt0)  =  Mx0)  —(to).  (2) 
dt  dx  dt 

Although  equation  (2)  is  most  useful  in  practice,  it  does  represent  an  unfor- 
tunate abuse  of  notation  in  that  the  symbol  /  is  used  to  denote  both  a  function 
of  x  and  one  of  t.  It  would  be  more  appropriate  to  define  a  new  function  y  by 
y(t)  =  (f  o  x)(t)  sothatdy/df  =  (df/dx)(dx/dt).  But  our  original  abuse  of  no- 
tation is  actually  a  convenient  one,  since  it  avoids  the  awkwardness  of  having  too 
many  variable  names  appearing  in  a  single  discussion.  In  the  name  of  simplicity, 
we  will  therefore  continue  to  commit  such  abuses  and  urge  you  to  do  likewise. 

The  formulas  in  equations  (1)  and  (2)  are  so  simple  that  little  more  needs 
to  be  said.  We  elaborate,  nonetheless,  because  this  will  prove  helpful  when  we 


1  44       Chapter  2  |  Differentiation  in  Several  Variables 


generalize  to  the  case  of  several  variables.  The  chain  rule  tells  us  the  following: 
To  understand  how  /  depends  on  t,  we  must  know  how  /  depends  on  the  "in- 
termediate variable"  x  and  how  this  intermediate  variable  depends  on  the  "final" 
independent  variable  t.  The  diagram  in  Figure  2.58  traces  the  hierarchy  of  the 
variable  dependences.  The  "paths"  indicate  the  derivatives  involved  in  the  chain 
rule  formula. 


Figure  2.58  The  chain  rule  for  functions  of  a  single  variable. 


Figure  2.59  The  composite  function  fox. 


The  Chain  Rule  in  Several  Variables 

Now  let's  go  a  step  further  and  assume  /:XcR2^RisaC'  function  of 
two  variables  and  x:  T  c  R  ->  R2  is  a  differentiable  vector-valued  function 
of  a  single  variable.  If  the  range  of  x  is  contained  in  X,  then  the  composite 
/ox:TcR^Ris  defined.  (See  Figure  2.59.)  It's  good  to  think  of  x  as 
describing  a  parametrized  curve  in  R2  and  /  as  a  sort  of  "temperature  func- 
tion" on  X.  The  composite  /  o  x  is  then  nothing  more  than  the  restriction  of  /  to 
the  curve  (i.e.,  the  function  that  measures  the  temperature  along  just  the  curve). 
The  question  is,  how  does  /  depend  on  t ?  We  claim  the  following: 


PROPOSITION  5.2  Suppose  x:  T  C  R  ->  R2  is  differentiable  at  t0  e  T,  and 
/:  X  C  R2  —>  R  is  differentiable  at  xo  =  x(?o)  =  (xq,  yo)  £  X,  where  T  and  X 
are  open  in  R  and  R2,  respectively,  and  range  x  is  contained  in  X.  If,  in  addition, 
/  is  of  class  C1,  then  fox:T^  R  is  differentiable  at  to  and 

df  df       dx  df  dy 

dt  dx       dt  dy  dt 


Before  we  prove  Proposition  5.2,  some  remarks  are  in  order.  First,  notice 
the  mixture  of  ordinary  and  partial  derivatives  appearing  in  the  formula  for  the 


2.5  |  The  Chain  Rule  145 


derivative.  These  terms  make  sense  if  we  contruct  an  appropriate  "variable  hier- 
archy" diagram,  as  shown  in  Figure  2.60.  At  the  intermediate  level,  /  depends 
on  two  variables,  x  and  y  (or,  equivalently,  on  the  vector  variable  x  =  (x ,  y)), 
so  partial  derivatives  are  in  order.  On  the  final  or  composite  level,  /  depends  on 
just  a  single  independent  variable  t  and,  hence,  the  use  of  the  ordinary  derivative 
df  /dt  is  warranted.  Second,  the  formula  in  Proposition  5.2  is  a  generalization  of 
equation  (2):  A  product  term  appears  for  each  of  the  two  intermediate  variables. 


Figure  2.61  The  graph  of  the 
function  x  of  Example  2. 


EXAMPLE  2  Suppose  f(x,  y)  =  (x  +  y2)/(2x2  +  1)  is  a  temperature  func- 
tion on  R2  and  x(?)  =  (2t,  t  +  1).  The  function  x  gives  parametric  equations 
for  a  line.  (See  Figure  2.61.)  Then 


(/ox)(0  =  /(x(0) 


2t  +  (t  +  iy    t2  +  At  +  i 


8r2  +  1  8?2  +  1 

is  the  temperature  function  along  the  line,  and  we  have 

df      A-  lAt  -  32t2 


dt 


(St2  +  l)2 


by  the  quotient  rule.  Thus,  all  the  hypotheses  of  Proposition  5.2  are  satisfied  and 
so  the  derivative  formula  must  hold.  Indeed,  we  have 


and 


Therefore, 


dx 
df 

dy 


At) 


\-2x2 


Axy2 


(2x2  +  l)2 
2y 


2x2  +  r 

dx  dy 
dt '  dt 


(2,  1). 


df  dx  df  dy 
dx  dt      dy  dt 


1  -  2x2  -  Axy2 
(2x2  +  l)2 

2(1  -  St2  -  8t(t 


2  + 


2v 


2jc2  +  1 


(8f2  +  l)2 


}f)  +  2(t 


1) 


8f2  +  1 


1  46       Chapter  2  |  Differentiation  in  Several  Variables 


after  substitution  of  2t  for  x  and  t  +  1  for  y.  Hence, 

df  dx     df  dy  _  2(2  -It-  \6t2) 
dx~dt  +  ~dy~dt  ~      (8?2  +  l)2  ' 
which  checks  with  our  previous  result  for  df/dt.  ♦ 

Proof  of  Proposition  5.2  Denote  the  composite  function  /  o  x  by  z-  We  want  to 
establish  a  formula  for  dz/dt  at  to-  Since  z  is  just  a  scalar- valued  function  of  one 
variable,  differentiability  and  the  existence  of  the  derivative  mean  the  same  thing. 
Thus,  we  consider 

dz  z(t)-z(t0) 

—(to)  =  km  — - —  , 

dt  t^t0     t  —  to 

and  see  if  this  limit  exists.  We  have 

dz,,  f(x(t),  y(t))  -  f(x(t0),  y(t0)) 

—(to)  =  bm  . 

dt  f-»-«o  t  —  to 

The  first  step  is  to  rewrite  the  numerator  of  the  limit  expression  by  subtracting 
and  adding  f(x0l  y)  and  to  apply  a  modicum  of  algebra.  Thus, 

dzr^    v   f(x'    -  /(*o»  y)  +  f(xo,  y)  -  f(x0,  y0) 

—(to)  =  bm  

dt  t->tQ  t  —  to 

,.    fix,  y)  -  f(x0,  y)          f(xo,  y)  -  f(x0,  y0) 
=  hm  h  bm  . 

t-*h         t  —  to  t-*t<,  t  —  to 

(Remember  that  x(to)  =  xo  =  (xo,  yo)-)  Now,  for  the  main  innovation  of  the  proof. 
We  apply  the  mean  value  theorem  to  the  partial  functions  of  /.  This  tells  us  that 
there  must  be  a  number  c  between  x0  and  x  and  another  number  d  between  y0 
and  y  such  that 


and 


Thus, 


f(x,  y)  -  f(x0,  y)  =  fx(c,  y)(x  -  x0) 
f(x0,  y)  -  f(x0,  yo)  =  fy(x0,  d)(y  -  y0). 


dz  x  —  xq  y  —  yo 

—(to)  =  hm  fx(c,  y)- — —  +  hm  fy(x0,  d)- — — 
dt  t->t0  t  —  to  ■  t  —  to 

x(t)-x(to)  ,,y(t)-y(to) 

=  hm  fx(c,  y) —  h  hm  fy(x0,  d) —  

t-*k  t  - 10         t^-to  t  -  to 

dx  dy 

=  fx(XQ,  yo)-J-(to)  +  fy(XQ,  yo)-T-(t0), 

dt  dt 
by  the  definition  of  the  derivatives 

dx  dy 
-(to)    and  -(t0) 

and  the  fact  that  fx(c,  y)  and  fy(xo,  d)  must  approach  fx(xo,  yo)  and  fy(xo,  yo), 
respectively,  as  t  approaches  to,  by  continuity  of  the  partials.  (Recall  that  /  was 
assumed  to  be  of  class  C 1 . )  This  completes  the  proof.  ■ 

Proposition  5.2  and  its  proof  are  easy  to  generalize  to  the  case  where  / 
is  a  function  of  n  variables  (i.e.,  /:XC  R"  —>  R)  and  x  :  T  C  R  — >•  R".  The 


2.5  |  The  Chain  Rule  147 


appropriate  chain  rule  formula  in  this  case  is 

df  ,  ,      df  ,   . dxi ,         df  ,    ,dx2,  s 

=  -^(x0)— ^0)  +  -^(x0)— i(r0)  + 
at  ax\       at  0x2  dt 


df  ,dx„ 

+  w~  (xo)-r-(fo)- 
3x„  a/ 


(3) 


Note  that  the  right  side  of  equation  (3)  can  also  be  written  by  using  matrix  notation 
so  that 


dj_ 
dt 


Co)  = 


9/ 
3xi 


(xo) 


Thus,  we  have  shown 


3/ 
3x2 


(xo) 


dx„ 


(xo) 


~dl 


dt 


(to) 
(to) 

(to) 


df 
dt 


(t0)  =  Df(x0)Dx(t0)  =  V/(x0)  -x'(/0), 


(4) 


where  we  use  x'(fo)  as  a  notational  alternative  to  Dx(fo)-  The  version  of  the  chain 
rule  given  in  formula  (4)  is  particularly  important  and  will  be  used  a  number  of 
times  in  our  subsequent  work. 


Let  us  consider  further  instances  of  composition  of  functions  of  many 
variables.  For  example,  suppose  X  is  open  in  R3,  T  is  open  in  R2,  and 
/:XcR3^R  and  x:  T  c  R2  R3  are  such  that  the  range  of  x  is  contained 
in  X.  Then  the  composite  /  os:rcR2->R  can  be  formed  as  shown  in 
Figure  2.62.  Note  that  the  range  of  x,  that  is,  x(7),  is  just  a  surface  in  R3,  so 
fox  can  be  thought  of  as  an  appropriate  "temperature  function"  restricted  to  this 
surface.  If  we  use  x  =  (x,  y,  z)  to  denote  the  vector  variable  in  R3  and  t  =  (s,t) 
for  the  vector  variable  in  R2,  then  we  can  write  a  plausible  chain  rule  formula 
from  an  appropriate  variable  hierarchy  diagram.  (See  Figure  2.63.)  Thus,  it  is 


X 


1  48       Chapter  2  |  Differentiation  in  Several  Variables 


Figure  2.63  The  chain  rule  for  /  o  x,  where  /:XCR!^R  and 
\:T  c  R2  ->-  R3. 


reasonable  to  expect  that  the  following  formulas  hold: 

df  =  dfdx  +  dfdy  dfdz 

ds       dx  ds      dy  ds      dz  ds 

and  (5) 

df  _  df  dx      df  dy      df  dz 
~dl  ~  dx~dl  +  ~dyJt  +  JzJt' 

(Again,  we  abuse  notation  by  writing  both  df/ds,  df/dt  and  df/dx,  df/dy, 
df/dz.)  Indeed,  when  /  is  a  function  of  x,  y,  and  z  of  class  C1,  formula  (3) 
with  n  =  3  applies  once  we  realize  that  dx/ds,  dx/dt,  etc.,  represent  ordinary 
differentiation  of  the  partial  functions  in  s  or  t. 

EXAMPLE  3  Suppose 

f(x,  y,  z)  =  x2  +  y2  +  z2    and    x(s,  t)  =  (s  cos  t,  est ,  s2  -  t2). 
Then  h(s,  t)  =  f  o  x(s,  t)  =  s2  cos2 1  +  e2st  +  (s2  -  t2)2,  so  that 

dh      d(f  ox)     ^       2  ~  ,  ,  7x 

=  2*  cos2  f  +  2fe2sr  +  4s(s2  -  t2) 


dt  dt 
We  also  have 


ds  ds 

=  -2s1  cos  t  sin  t  +  2seIst  -  4t(sz  -  tl). 


dh      d(fox)  2  ,,  2 


and 


9/     -         3/     -         3/  - 
—  =  2x,       —  =  2y,       —  =  2z 
dx  dy  dz 


dx  dx 

—  =  cosr,  —  =  —  jsin/, 
ds  dt 

t  st  dy  st 

—  =  te  ,  —  =  se  , 

ds  dt 

3z  dz 

—  =2s,  —  =  -It. 
ds  dt 


2.5  |  The  Chain  Rule  149 


Hence,  we  compute 

df  =  d(f  o  x)  =  df  dx  |  df  dy  |  df  dz 
ds  ds  dx  ds      dy  ds      dz  ds 

=  2x(cos  0  +  2y(test)  +  2z(2s) 

=  2s  cos  r(cos  0  +  2est(test)  +  2{s2  -  t2)(2s) 

=  2s  cos2  t  +  2te2s1  +  4s(s2  -  t2), 

just  as  we  saw  earlier.  We  leave  it  to  you  to  use  the  chain  rule  to  calculate  df/dt 
in  a  similar  manner.  ♦ 

Of  course,  there  is  no  need  for  us  to  stop  here.  Suppose  we  have  an  open  set 
X  in  R'",  an  open  set  T  in  R",  and  functions  /:  X  —>  R  and  x:  T  — »•  R'"  such 
that  A  =  /  o  x:  I  ^>  R  can  be  defined.  If  /  is  of  class  C1  and  x  is  differentiate, 
then,  from  the  previous  remarks,  h  must  also  be  differentiate  and,  moreover, 

dh       df  dxi      df  3x2  3/  dxm 

dtj       dx\  dtj       dx2  dtj  dxm  dtj 


=  E 


7  =  1,2, 


df  dxk 
k=l  dxk  dtj 

Since  the  component  functions  of  a  vector- valued  function  are  just  scalar- valued 
functions,  we  can  say  even  more.  Suppose  f:  X  c  Rm  ->  RP  and  x:  T  c  R" 
R"!  are  such  that  h  =  f  o  x:  T  c  R"  — >•  R''  can  be  defined.  (As  always,  we  assume 
that  X  is  open  in  Rm  and  T  is  open  in  R" .)  See  Figure  2.64  for  a  representation  of 
the  situation.  If  f  is  of  class  C 1  and  x  is  differentiable,  then  the  composite  h  =  f  o  x 
is  differentiable  and  the  following  general  formula  holds: 


dhi 
3^ 


df  dxk 
dxk  dtJ 


1,2, 


7  =  1,2, 


,  n. 


(6) 


The  plausibility  of  formula  (6)  is  immediate,  given  the  variable  hierarchy  diagram 
shown  in  Figure  2.65. 


X 


x(T) 


R"  Rm  Rp 

Figure  2.64  The  composite  fo  x  where  f :  X  C  R"'  ^  R''  and  x:  T  C  R"  Rm. 

Now  comes  the  real  "magic."  Recall  that  if  A  is  a  p  x  m  matrix  and  B  is  an 
m  x  n  matrix,  then  the  product  matrix  C  =  AB  is  defined  and  is  a  /?  x  «  matrix. 
Moreover,  the  i/th  entry  of  C  is  given  by 


kj- 


k=l 


Chapter  2  |  Differentiation  in  Several  Variables 


If  we  recall  that  the  ijth  entry  of  the  matrix  Dh(t)  is  dhi/dtj,  and  similarly  for 
Z)f(x)  and  Dx(t),  then  we  see  that  formula  (6)  expresses  nothing  more  than  the 
following  equation  of  matrices: 

Dh(t)  =  D(f  o  x)(t)  =  Df(x)Dx(t).  (7) 

The  similarity  between  formulas  (7)  and  (1)  is  striking.  One  of  the  reasons 
(perhaps  the  principal  reason)  for  defining  matrix  multiplication  as  we  have  is 
precisely  so  that  the  chain  rule  in  several  variables  can  have  the  elegant  appearance 
that  it  has  in  formula  (7). 

EXAMPLE  4   Suppose     f:R3  ->  R2     is     given     by     f(xi,x2,x3)  = 
(jci  —  x2,  xix2x3)  and  x:  R2  ->  R3  is  given  by  x(*i,  t2)  =  (ht2,  t2 
f  o  x:  R2  ->  R2  is  given  by  (f  o  x)(*i ,  t2)  =  (tit2  -  tf,  f3f23),  so  that 


ls  4).  Then 


D(fox)(t) 

On  the  other  hand 

Df(X)  : 

so  that  the  product  matrix  is 
Df(x)Dx(t) 


t2  -  2fi 
3t24 


h 

3ty2 


1       -1  0 

X2Xj,        X\Xt,  X\X2 


h  ~  2*i 


and  Dx(t) 


h  *i 
2*!  0 
0  2*2 


X2X-x,t2  +  2x  1X3*1        X2*3fi  +  2x\X2t2 


*2 

t2t3 

'1  '2 


2*1 
2f2 


,3 

1  '2 


2t\t\ 


after  substituting  forxi,  x2,  andx3.  Thus,  D(f  o  x)(t)  =  Df(x)Dx(t),  as  expected. 

Alternatively,  we  may  use  the  variable  hierarchy  diagram  shown  in  Figure  2.66 
and  compute  any  individual  partial  derivative  we  may  desire.  For  example, 

dfl  _  df^dx^  dfidx^  df2  9x3 
3*i      9xi  3*i      3x2  3?i      3x3  3*i 


2.5  |  The  Chain  Rule  151 


Figure  2.66  The  variable  hierarchy  diagram  for  Example  4. 

by  formula  (6).  Then  by  abuse  of  notation, 

3/2 


=  (*2*3)(fc)  +  (*l*3)(2fi)  +  (XIX2)(0) 


=  (t2t22)(t2)  +  (ht2)(t22)(2h) 


which  is  indeed  the  (2,1)  entry  of  the  matrix  product.  ♦ 

At  last  we  state  the  most  general  version  of  the  chain  rule  from  a  technical 
standpoint;  a  proof  may  be  found  in  the  addendum  to  this  section. 


THEOREM  5.3  (The  chain  rule)  Suppose  X  c  R"!  and  T  c  R"  are  open  and 
f:  X  — >  Kp  and  x:  T  — »•  R'"  are  defined  so  that  range  x  c  X.  If  x  is  differen- 
tiable  at  t0  6  T  and  f  is  differentiable  at  x0  =  x(t0),  then  the  composite  f  o  x  is 
differentiable  at  to,  and  we  have 


D(fox)(t0)=  Df(x0)Dx(t0). 


The  advantage  of  Theorem  5.3  over  the  earlier  versions  of  the  chain  rule  we 
have  been  discussing  is  that  it  requires  f  only  to  be  differentiable  at  the  point  in 
question,  not  to  be  of  class  C1.  Note  that,  of  course,  Theorem  5.3  includes  all 
the  special  cases  of  the  chain  rule  we  have  previously  discussed.  In  particular, 
Theorem  5.3  includes  the  important  case  of  formula  (4). 

EXAMPLE  5  Let  f:  R2  ->  R2  be  defined  by  f(x,  y)  =  (x  -  2y  +  7,  3xy2). 
Suppose  that  g:  R3  ->  R2  is  differentiable  at  (0,  0,  0)  and  we  know  that 
g(0,  0,  0)  =  (-2,  1)  and 

2  4  5  " 
Dg(0,0,0)=    _i  o  1  ■ 

We  use  this  information  to  determine  D(f  o  g)(0,  0,  0). 

First,  note  that  Theorem  5.3  tells  us  that  fog  must  be  differentiable  at  (0,  0,  0) 
and,  second,  that 


D(f  o  g)(0,  0,  0)  =  Df  (g0,  0,  0))  Dg(0,  0,  0)  =  Df(-2,  l)Dg(0,  0,  0). 


Chapter  2  |  Differentiation  in  Several  Variables 


Since  we  know  f  completely,  it  is  easy  to  compute  that 
Df(x,y)  = 

Thus, 

D(fog)(0,0,0)=  3 

We  remark  that  we  needed  the  full  strength  of  Theorem  5.3,  as  we  do  not  know 
anything  about  the  differentiability  of  g  other  than  at  the  point  (0,  0,  0).  ♦ 


1  -2 

3y2  6xy 


so  that    Df(-2,  1) 


-2 
■12 


2  4  5 
-1   0  1 


4  3 
12  3 


EXAMPLE  6  (Polar/rectangular  conversions)  Recall  that  in  §1.7  we  pro- 
vided the  basic  equations  relating  polar  and  rectangular  coordinates: 

x  =  r  cos  9 
y  =  r  sin  9 

Now  suppose  you  have  an  equation  defining  a  quantity  w  as  a  function  of  x  and 
y;  that  is, 

w  =  f{x,  y). 

Then,  of  course,  w  may  just  as  well  be  regarded  as  a  function  of  r  and  9  by 
susbtituting  r  cos  9  for  x  and  r  sin  9  for  y  •  That  is, 


g(r,  0)  =  f(x(r,  9),  y(r,  9)). 


Our  question  is  as  follows:  Assuming  all  functions  involved  are  differentiable, 
how  are  the  partial  derivatives  dw/dr,  dw/d9  related  to  dw/dx,  dw/dy? 

In  the  situation  just  described,  we  have  w  =  g(r,  9)  =  (/  o  x)(r,  9),  so  that 
the  chain  rule  implies 


Dg(r,9)  =  Df(x,y)Dx(r,9). 


Therefore, 


dg 
8r 


9g 
36 


dx 


df 

dx 


df 

dy 

df 
dy 


dx 

dx 

W 

dy 

dy 

dr 

d9  _ 

cos  9 

—r  sin 

sin  9 

r  cos 

By  extracting  entries,  we  see  that  the  various  partial  derivatives  of  w  are  related 
by  the  following  formulas: 


3w 
"97 
dw 
~d~9 


cost 


dw  dw 

 h  sm9  — 

dx  dy 


dw  dw 

-r  smw  h  r  zos9  — 

dx  dy 


(8) 


The  significance  of  (8)  is  that  it  provides  us  with  a  relation  of  differential 
operators: 


2.5  |  The  Chain  Rule  153 


a  a 

—  =  cost?  — 
dr  dx 


sinO 


dy 


d  a  a 

—  =  —r  sin  6  h  r  cos  6  — 

dd  dx  dy 


(9) 


The  appropriate  interpretation  for  (9)  is  the  following:  Differentiation  with 
respect  to  the  polar  coordinate  r  is  the  same  as  a  certain  combination  of  differen- 
tiation with  respect  to  both  Cartesian  coordinates  x  and  y  (namely,  the  combina- 
tion cos  0  d/dx  +  sm6  d/dy).  A  similar  comment  applies  to  differentiation  with 
respect  to  the  polar  coordinate  0 .  Note  that,  when  r  ^  0,  we  can  solve  algebraically 
for  d/dx  and  d/dy  in  (9),  obtaining 


a 

—  =  cosf 
dx 

sin  6 

d 

dr 

/• 

dQ 

a 

—  =  sinfi 

dy 

a 

Yr  + 

cos  6 

a 

r 

d9 

(10) 


We  will  have  occasion  to  use  the  relations  in  (9)  and  (10),  and  the  method  of 
their  derivation,  later  in  this  text.  ♦ 

Addendum:  Proof  of  Theorem  5.3   


We  begin  by  noting  that  the  derivative  matrices  £>f(x0)  and  Dx(t0)  both  exist 
because  f  is  assumed  to  be  differentiable  at  x0  and  x  is  assumed  to  be  differentiable 
at  to.  Thus,  the  product  matrix  Df(xo)£>x(to)  exists.  We  need  to  show  that  the  limit 
in  Definition  3.8  is  satisfied  by  this  product  matrix,  that  is,  that 

||(fox)(t)-[(fox)(t0)  +  Df(xo)Dx(t0)(t-to)]|| 

hm  =  0.  (11) 

||t -toll 

In  view  of  the  uniqueness  of  the  derivative  matrix,  it  then  automatically  follows 
that  f  o  x  is  differentiable  at  to  and  that  Z)f(xo)£>x(to)  =  D(f  o  x)(to).  Thus,  we 
entirely  concern  ourselves  with  establishing  the  limit  (11)  above. 
Consider  the  numerator  of  (1 1).  First,  we  rewrite 

(f  o  x)(t)  -  [(f  ox)(to)  +  Df(x0)Dx(t0)(t  -  t0)] 

=  (f  o  x)(t)  -  (f  o  x)(t0)  -  £>f(x0)(x(t)  -  x(f0)) 

+  Df(x0)(x(t)  -  x(f0))  -  Df(xo)Dx(t0)(t  -  t0). 

Then  we  use  the  triangle  inequality: 

||  (f  o  x)(t)  -  [(f  o  x)(t0)  +  £>f(x0)Dx(to)(t  -  t0)]  || 

<  || (f  o  x)(t)  -  (f  o  x)(t0)  -  Df(x0)(x(t)  -  x(t0))\\ 

+  ||£»f(x0)(x(t)  -  x(f0))  -  Df(x0)Dx(t0)(t  -  t0)  || 
=  ||  (f  o  x)(t)  -  (f  o  x)(t0)  -  Df(x0)(x(t)  -  x(?0))|| 

+  ||Df(xo)  [(x(t)  -  x(f0))  -  Dx(t0)(t  -  t0)]  || . 


1  54       Chapter  2  |  Differentiation  in  Several  Variables 


By  inequality  (9)  in  the  proof  of  Theorem  3.9,  there  is  a  constant  K  such  that,  for 
any  vector  h  e  R",  ||Df(x0)h||  <  A"||h||.  Thus, 

|| (fo  x)(t)  -  (f  o  x)(t0)  -  £>f(x0)Dx(t0)(t  -  t0) || 

<  || (f  o  x)(t)  -  (f  o  x)(t0)  -  Df(x0)(x(t)  -  x(t0))\\  (12) 
+  K||x(t)-x(fo)-JDx(t0Xt-to)||. 

To  establish  the  limit  (11)  formally,  we  must  show  that  given  any  e  >  0,  we 
may  find  a  S  >  0  such  that  if  0  <  ||t  —  toll  <  8,  then 

||  (f  o  x)(t)  -  [(f  o  x)(t0)  +  £>f(xo)Z)x(to)(t  -  t0)]  ||  <f 

lit -toll 

Consider  the  first  term  of  the  right  side  of  (12).  Using  the  differentiability  of  x  at 
t0  and  inequality  (1 1)  in  the  proof  of  Theorem  3.9,  we  can  find  some  <50  >  0  and 
a  constant  A"o  such  that  if  0  <  ||t  —  toll  <  So,  then 

||x(t)  -  x(t0)||  <  K0  ||t  -to||. 

By  the  differentiability  of  f  at  xo,  given  any  e\  >  0,  we  may  find  some  Si  >  0 
such  that  if  0  <  ||x  —  Xo||  <  <$i,  then 

||f(x)-[f(xo)  +  Df(x0)(x-xo)]  || 

  <  ei- 


l|x-xo|| 

Setei  =  e/(2A:0).  Withx  =  x(t),x0  =  x(t0),  we  have  that  if  both  0  <  ||t-  t0||  < 
So  and  0  <  ||t  —  t0||  <  8\/Kq,  then 

||x(t)  -  x(t0)||  <  Ko  lit -toll  <<5i. 

Hence, 

||(f  o  x)(t)  -  (f  o  x)(t0)  -  Z)f(x0)(x(t)  -  x(t0))||  <  ei||x(t)  -  x(t0)|| 

<  ei^ollt-  toll  =  .  (13) 

Now  look  at  the  second  term  of  the  right  side  of  (12).  Since  x  is  differentiable 
at  t0,  given  any  €2  >  0,  we  may  find  some  S2  >  0  such  that  if  0  <  ||t  —  t0 1|  <  S2, 
then 

||x(t)  -  [x(t0)  +  £>x(t0)(t  -  t0)]  || 
lit -toll 

Set  e2  =  e/(2K).  Then,  for  0  <  ||t  -  t0||  <  S2,  we  have 

||x(t)-[x(to)  +  Dx(to)(t-t0)]||  <  2^llt-t0||.  (14) 

Finally,  let  S  be  the  smallest  of  <5o,  <5i/A"o,  and  82-  Then,  for  0  <  ||t  —  to ||  <  8,  we 
have  that  both  the  inequalities  (13)  and  (14)  hold  and  thus  (12)  becomes 

||  (fo  x)(t)  -  (f  o  x)(t0)  -  Df(x0)£»x(t0)(t  -  t0)  || 

<  €-\\t  -  toll  +^(^11*- toll) 

=  e||t- toll- 
Hence, 

||(f  o  x)(t)  -  (f  o  x)(t0)  -  £)f(xo)Dx(to)(t  -  t0)" 
lit -toll 

as  desired. 


2.5  I  Exercises  155 


2.5  Exercises 


1.  If  f(x,  y,  z)  =  x2  —  y3  +  xyz,  and  x  =  6t  +  7,  y  = 
sin2r,  z  =  t2,  verify  the  chain  rule  by  finding  df /dt 
in  two  different  ways. 


months  from  now  will  be 


71 1 


s2  +  t2, 


find 


2.  If  f(x,  y)  =  sin(;ty)  and x  =  s  +  t,  y 
df/ds  and  df/dt  in  two  ways: 

(a)  by  substitution. 

(b)  by  means  of  the  chain  rule. 

3.  Suppose  that  a  bird  flies  along  the  helical  curve 
x  =  2  cos  t,  y  =  2  sin  t,z  =  3f .  The  bird  suddenly  en- 
counters a  weather  front  so  that  the  barometric  pres- 
sure is  varying  rather  wildly  from  point  to  point  as 
P(x,  y,  z)  =  6x2z/y  atm. 

(a)  Use  the  chain  rule  to  determine  how  the  pressure 
is  changing  at  t  =  rc/4  min. 

(b)  Check  your  result  in  part  (a)  by  direct  substitution. 

(c)  What  is  the  approximate  pressure  at  t  =  rc/4  + 
0.01  min? 


4.  Suppose  that  z 


+  y  ,  where  x  =  st  and  y  is  a 


function  ofs  and  t .  Suppose  further  that  when  (s,  t)  = 

dz 

(2,  1),  dy/dt  =  0.  Determine  —(2,  1). 

5.  You  are  the  proud  new  owner  of  an  Acme  Deluxe  Bread 
Kneading  Machine,  which  you  are  using  for  the  first 
time  today.  Suppose  that  at  noon  the  dimensions  of  your 
(nearly  rectangular)  loaf  of  bread  dough  are  L  =  7  in 
(length),  W  =  5  in  (width),  and  H  =  4  in  (height).  At 
that  time,  you  place  the  loaf  in  the  machine  for  knead- 
ing and  the  machine  begins  by  stretching  the  loaf's 
length  at  an  initial  rate  of  0.75  in/min,  punching  down 
the  loaf's  height  at  a  rate  of  1  in/min,  and  increasing 
the  loaf's  width  at  a  rate  of  0.5  in/min.  What  is  the  rate 
of  change  of  the  volume  of  the  loaf  when  the  machine 
starts?  Is  the  dough  increasing  or  decreasing  in  size  at 
that  moment? 

6.  A  rectangular  stick  of  butter  is  placed  in  the  microwave 
oven  to  melt.  When  the  butter's  length  is  6  in  and  its 
square  cross  section  measures  1 .5  in  on  a  side,  its  length 
is  decreasing  at  a  rate  of  0.25  in/min  and  its  cross- 
sectional  edge  is  decreasing  at  a  rate  of  0.125  in/min. 
How  fast  is  the  butter  melting  (i.e.,  at  what  rate  is  the 
solid  volume  of  butter  turning  to  liquid)  at  that  instant? 

7.  Suppose  that  the  following  function  is  used  to  model 
the  monthly  demand  for  bicycles: 

P(x,  y)  =  200  +  20V0.1*  +  10  -  lltfy. 

In  this  formula,  x  represents  the  price  (in  dollars 
per  gallon)  of  automobile  gasoline  and  y  repre- 
sents the  selling  price  (in  dollars)  of  each  bicycle. 
Furthermore,  suppose  that  the  price  of  gasoline  t 


x  =  1  +  O.lf  —  cos  ■ 
b 

and  the  price  of  each  bicycle  will  be 

TCt 

y  =  200  +  2t  sin — . 

6 

At  what  rate  will  the  monthly  demand  for  bicycles  be 
changing  six  months  from  now? 

8.  The  Centers  for  Disease  Control  and  Prevention  pro- 
vides information  on  the  body  mass  index  (BMI) 
to  give  a  more  meaningful  assessment  of  a  person's 
weight.  The  BMI  is  given  by  the  formula 


BMI : 


10,000u; 


where  w  is  an  individual's  mass  in  kilograms  and  h 
the  person's  height  in  centimeters.  While  monitoring  a 
child's  growth,  you  estimate  that  at  the  time  he  turned 
1 0  years  old,  his  height  showed  a  growth  rate  of  0.6  cm 
per  month.  At  the  same  time,  his  mass  showed  a  growth 
rate  of  0.4  kg  per  month.  Suppose  that  he  was  140  cm 
tall  and  weighed  33  kg  on  his  tenth  birthday. 

(a)  At  what  rate  is  his  BMI  changing  on  his  tenth 
birthday? 

(b)  The  BMI  of  a  typical  10-year-old  male  increases 
at  an  average  rate  of  0.04  BMI  points  per  month. 
Should  you  be  concerned  about  the  child's  weight 
gain? 

9.  A  cement  mixer  is  pouring  concrete  in  a  conical  pile. 
At  the  time  when  the  height  and  base  radius  of  the 
concrete  cone  are,  respectively,  30  cm  and  12  cm,  the 
rate  at  which  the  height  is  increasing  is  1  cm/min  and 
the  rate  at  which  the  volume  of  cement  in  the  pile  is 
increasing  is  320  cm3/min.  At  that  moment,  how  fast 
is  the  radius  of  the  cone  changing? 

1 0.  A  clarinetist  is  playing  the  glissando  at  the  beginning  of 
Rhapsody  in  Blue,  while  Hermione  (who  arrived  late) 
is  walking  toward  her  seat.  If  the  (changing)  frequency 
of  the  note  is  /  and  Hermione  is  moving  toward  the 
clarinetist  at  speed  v,  then  she  actually  hears  the  fre- 
quency (j>  given  by 


c  +  v 


f, 


where  c  is  the  (constant)  speed  of  sound  in  air,  about 
330  m/sec.  At  this  particular  moment,  the  frequency  is 
/  =  440  Hz  and  is  increasing  at  a  rate  of  100  Hz  per 
second.  At  that  same  moment,  Hermione  is  moving 
toward  the  clarinetist  at  4  m/sec  and  decelerating  at 


Chapter  2  |  Differentiation  in  Several  Variables 


2  m/sec2.  What  is  the  perceived  frequency  cf>  she 
hears  at  that  moment?  How  fast  is  it  changing?  Does 
Hermione  hear  the  clarinet's  note  becoming  higher  or 
lower? 

11.  Suppose  z  =  f(x,y)  has  continuous  partial  deriva- 
tives. Let  x  =  e'cos8,  y  =  e'smd.  Show  that 
then 


+ 


-2r 


+ 


dz. 


X  z 


1 8.  Suppose  that  ui  =  g  I  — ,  —  I  is  a  differentiable  func- 

\y  yJ 

tion  of  u  =  x/y  and  v  =  z/y.  Show  then  that 


dw       dw  dw 

xir  +  yir  +  zir=0- 

ox         ay  dz 


In  Exercises  19-27,  calculate  D(io  g)  in  two  ways:  (a)  by  first 
evaluating  fog  and  (b)  by  using  the  chain  rule  and  the  deriva- 
tive matrices  Df  and  Dg. 


12.  Suppose  that  z  =  f(x,  y)  has  continuous  partial 
derivatives.  Let  x  =  2uv  and  y  =  u2  +  v2 .  Show  that 
then 


dz  dz 


2x 


du  dv 
13.  If  w  =  g(u2 


+ 


,     dz  dz 

+  4v  . 

dx  3y 


u2)  has  continuous  partial 


derivatives  with  respect  to  x  =  u2  —  v2  and  y  =  v2  — 
u2,  show  that 

dw  dw 

v  h  u  —  =0. 

du  dv 

14.  Suppose  that  z  =  f(x  +  y,  x  —  y)  has  continuous  par- 
tial derivatives  with  respect  to  u  =  x  +  y  and  v  = 
x  —  y.  Show  that 


dz 

dz 

(du) 

(dz 

dx 

dy  " 

\dv 

15.  If  w  =  f 


xy 


xy 


x2  +  y 


is  a  differentiable  function  of 


- ,  show  that 


x2  +  y2 

dw  dw 

x  h  y        =  0. 

dx  dy 

(x2-y2\ 

16.  If  w  =  f    —z  r-    is  a  differentiable  function  of 

\x2  +  y2J 

x2  -  y2 
u  =  —  ,  show  that  then 


xl  +  y1 


dw  dw 

x  h  y         =  0. 

dx  dy 


I  y  —  x  z  —  x  \ 
1 7.  Suppose  w  =  f  [   ,  )  is  a  differentiable 

V  xy  xz 

y  —  x            z  —  x 
function  of  u  =   and  v  =   .  Show  then  that 


xy 


xz 


r.  dw        i  dw       i  dw 

x  ir + y  IT + z  IT  =  °- 

dx  dy  dz 


19.  f(x)  =  (3x5,  e2x),  g(s,  t)  =  s-  It 

20.  f(x)  =  (x2,  cosSx,  Inx),  g(s,  t,u)  =  s  +  t2  +  i 

21.  f(x,  y)  =  yex,  g(s,  t)  =  (s  -  t,  s  +  t) 


22.  f(x,  y) 

23.  f(x,  y)  = 


3y2,g(s,t)  =  (st,s  +  t2) 


(xv-^  +  v3),g(,,r)=(^2r) 


y  x 
x'  y 

24.  f(x,  y,  z)  =  (x2y  +  y2z,  xyz,  ez), 
g(t)  =  (f  -  2,  3f  +  7,  f3) 

25.  f(x,  y)  =  (xy2,  x2y,x3  +  y3),  g(f)  =  (sinf,  e') 

26.  f(x,  y)  =  (x2  -  y,  y/x,  ey),      g(s,  t,  u)  =  (s  +  It  + 
3m, stu) 

27.  f(x,  y,  z)  =  (x  +  y  +  z,  x3  -  evz), 
g(s,  t,  u)  =  (st,  tu,  su) 

28.  Let    g:  R3  — >•  R2    be    a    differentiable  function 
such  that  g(l,-l,3)  =  (2, 5)  and  Dg(l,-1,3)  = 

1    -1  0 


0  7 


Suppose  that  f:  R  — >  R    is  de- 


fined by  f(x,  v)  =  (2xy,  3x  —  y  +  5).  What  is 
D(fog)(l,-l,3)? 

29.  Let  g:R2^R2  and  f:R2^R2  be  differentiable 
functions  such  that  g(0, 0)  =  (1, 2),  g(l,2)  = 
(3,5),  f(0,  0)  =  (3,  5),  f(4,  1)  =  (1,2),  Z)g(0,0)  = 


1  0 
-1  4 


1  1 

3  5 


Dg{\,2): 

Df(4,  1)  = 


2  3 
5  7 


£>f(3,  5)  = 


1  2 
1  3 


(a)  Calculate  D(fog)(  1,2). 

(b)  Calculate  D(gof)(4,  1). 

30.  Let  z  =  f(x,y),  where  /  has  continuous  partial 
derivatives.  If  we  make  the  standard  polar/rectangular 
substitution  x  =  r  cos  0,  y  =  r  sin  9,  show  that 


+ 


+ 


1 


.30. 


2.5  I  Exercises  157 


31.  (a)  Use  the  methods  of  Example  6  and  formula  (10) 
in  this  section  to  determine  d2/dx2  and  d2/dy2 
in  terms  of  the  polar  partial  differential  operators 
d2/dr2,  d2/dd2,  d2/dr  d6,  d/dr,  and  d/dd.  (Hint: 
You  will  need  to  use  the  product  rule.) 

(b)  Use  part  (a)  to  show  that  the  Laplacian  operator 

d2/dx2  +  d2/dy2  is  given  in  polar  coordinates  by 
the  formula 


il  il 

Sx2  +  dy2 


dr2 


+ 


1  3 
r  dr 


+ 


l 

3<92' 


32.  Show  that  the  Laplacian  operator  d2/dx2  +  d2/dy2  + 
d2/dz2  in  three  dimensions  is  given  in  cylindrical  co- 
ordinates by  the  formula 


32  d2  d2 
dx2      By2  dz2 


d2       13  1 


■dr- 


r  dr 


+ 


r2  d62  dz 


33.  In  this  problem,  you  will  determine  the  formula  for  the 
Laplacian  operator  in  spherical  coordinates. 

(a)  First,  note  that  the  cylindrical/spherical  conver- 
sions given  by  formula  (6)  of  §1.7  express  the 
cylindrical  coordinates  z  and  r  in  terms  of  the 
spherical  coordinates  p  and  <p  by  equations  of  pre- 
cisely the  same  form  as  those  that  express  x  and 
y  in  terms  of  the  polar  coordinates  r  and  9.  Use 
this  fact  to  write  3/3r  in  terms  of  3/3p  and  d/d<p. 
(Also  see  formula  (10)  of  this  section.) 

(b)  Use  the  ideas  and  result  of  part  (a)  to  establish  the 
following  formula: 

32  32  32 
d^2  +  dy~2  +  dz2 


+ 


1  32 


3p2     p2  dip2 


+ 


pz  sin"  cp 


so2 


2  3      cot  go 

p  dp  p2 


dcp 


34.  Suppose  that  y  is  defined  implicitly  as  a  function  y(x) 
by  an  equation  of  the  form 

F(x,  y)  =  0. 

(For  example,  the  equation  x3  —  y2  =  0  defines  y  as 
two  functions  of  x,  namely,  y  =  x3/2  and  y  =  — x3/2. 
The  equation  sin(xy)  —  x2y1  +  ey  =  0,  on  the  other 
hand,  cannot  readily  be  solved  for  y  in  terms  of  x.  See 
the  end  of  §2.6  for  more  about  implicit  functions.) 
(a)  Show  that  if  F  and  y(x)  are  both  assumed  to  be 
differentiable  functions,  then 

dy_  _  Fx(x,  y) 
dx         Fy(x,  y) 

provided  Fy(x,  y)  #  0. 


(b)  Use  the  result  of  part  (a)  to  find  dy /dx  when  y 
is  defined  implicitly  in  terms  of  x  by  the  equa- 
tion x3  —  y2  =  0.  Check  your  result  by  explicitly 
solving  for  y  and  differentiating. 

35.  Find  dy /dx  when  y  is  defined  implicitly  by  the  equa- 
tion sin(xy)  —  x2y7  +  ey  =  0.  (See  Exercise  34.) 

36.  Suppose  that  you  are  given  an  equation  of  the  form 

F(x,y,z)  =  Q, 

for    example,    something    like    x3z  +  y  cosz  + 
(sin  y)/z  =  0.  Then  we  may  consider  z  to  be  defined 
implicitly  as  a  function  z(x,  y). 
(a)  Use  the  chain  rule  to  show  that  if  F  and  z(x ,  y)  are 
both  assumed  to  be  differentiable,  then 


3z 
dx 


Fx(x,  y,  z) 
Fz(x,y,z)' 


dz 
dy 


Fy(x,y,z) 
Fz(x,  y,  z) 


(b)  Use  part  (a)  to  find  3z/3x  and  3z/3y  where  z  is 
given  by  the  equation  xyz  =  2.  Check  your  result 
by  explicitly  solving  for  z  and  then  calculating  the 
partial  derivatives. 

37.  Find  dz/dx  and  3z/3y,  where  z  is  given  implicitly  by 
the  equation 


,                   sin  y 
x  z  +  y  cos  z  H  =  0. 


(See  Exercise  36.) 
38.  Let 


f(x,y) 


2 

x  y 
x2  +  y2 

0 


if  (x,  y)  #  (0,  0) 
if(x,y)  =  (0,0) 


(a)  Use  the  definition  of  the  partial  derivative  to  find 
fx(0,  0)  and  fy(0,  0). 

(b)  Let  a  be  a  nonzero  constant  and  let  x(f)  = 
(t,  at).  Show  that  /  o  x  is  differentiable,  and  find 
D(f  o  x)(0)  directly. 

(c)  Calculate  Df(0,  O)Z)x(O).  How  can  you  reconcile 
your  answer  with  your  answer  in  part  (b)  and  the 
chain  rule? 

Let  w  =  f(x,  y,  z)  be  a  differentiable  function  of  x,  y,  and 
Z.  For  example,  suppose  that  w  =  x  +  2y  +  z.  Regarding  the 
variables  x,  y,  and  z  as  independent,  we  have  dw/dx  =  1  and 
dw/dy  =  2.  But  now  suppose  that  z  =  xy.  Then  x,  y,  and  z 
are  not  all  independent  and,  by  substitution,  we  have  that  w  = 
X  +  2y  +  xy  so  that  dw/dx  =  1  +  y  and  dw/dy  =  2  +  x.  To 
overcome  the  apparent  ambiguity  in  the  notation  for  partial 
derivatives,  it  is  customary  to  indicate  the  complete  set  of  in- 
dependent variables  by  writing  additional  subscripts  beside 


Chapter  2  |  Differentiation  in  Several  Variables 


the  partial  derivative.  Thus, 


/9uA 


would  signify  the  partial  derivative  of  w  with  respect  to  x, 
while  holding  both  y  and  z  constant.  Hence,  x,  y,  and  z  are  the 
complete  set  of  independent  variables  in  this  case.  On  the  other- 
hand,  we  would  use  (dw/dx)y  to  indicate  thatx  andy  alone  are 
the  independent  variables.  In  the  case  that  w  =  x  +  2y  +  z, 
this  notation  gives 


9  w 
dx 


(  dw  \  (  dw 
1,     I          I     =  2,     and  I   

V  dy  Jx,z  V  dz 


1. 


If  Z  =  xy,  then  we  also  have 
dw\ 

  =  1  +  V,  and 

dx  /„ 


dw 
97 


2  +  x. 


In  this  way,  the  ambiguity  of  notation  can  be  avoided.  Use  this 
notation  in  Exercises  39-45. 


39.  Let  w  =  x  +  ly  —  lOz  and  z  =  x2  +  y2. 

dw\        /  dw\        /  dw\        f  dw 

97j,,,;  lay  A*'  \9z)x,y'  la* 

'  dw 


(a)  Find 


and 


dy 


(b)  Relate  (dw/dx)yz  and  (dw/dx)y  by  using  the 
chain  rule. 

40.  Repeat  Exercise  39  where  w  =  x3  +  y2  +  z3  and  z  = 

2x  —  3y. 


41.  Suppose  s  =  x2y  +  xzw  —  z2  and  xyw  —  v3z  +  xz 
=  0.  Find 


and 


ds 


42.  Let  U  =  F(P ',  V,  T)  denote  the  internal  energy  of  a 
gas.  Suppose  the  gas  obeys  the  ideal  gas  law  P  V  =  kT, 
where  &  is  a  constant. 
,'dU" 

(a)  Find 


(b)  Find 


(c)  Find 


3T 

dU 
~df 

dU 
~d~P 


43 .  Show  that  if  x ,  y ,  z  are  related  implicitly  by  an  equation 
of  the  form  F(x,  y,  z)  =  0,  then 


a*\  (dy\  (dz 

dy)z  \dz)x  \dxjv 


■1. 


This  relation  is  used  in  thermodynamics.  (Hint:  Use 
Exercise  36.) 

44.  The  ideal  gas  law  PV  =  kT,  where  £  is  a  constant, 
relates  the  pressure  P,  temperature  T,  and  volume  V 
of  a  gas.  Verify  the  result  of  Exercise  43  for  the  ideal 
gas  law  equation. 

45.  Verify  the  result  of  Exercise  43  for  the  ellipsoid 


ax2  +  by2  +  cz2  =  d 


where  a,  b,  c,  and  d  are  constants. 


2.6   Directional  Derivatives  and  the  Gradient 

In  this  section,  we  will  consider  some  of  the  key  geometric  properties  of  the 
gradient  vector 

v/  =  (i,i  *L) 

\dx\    dx2  dx„J 

of  a  scalar-valued  function  of  n  variables.  In  what  follows,  n  will  usually  be  2 
or  3. 

The  Directional  Derivative  

Let  f(x ,  y)  be  a  scalar- valued  function  of  two  variables.  In  §2.3,  we  understood  the 
partial  derivative  j^{a,b)  as  the  slope,  at  the  point  (a,  b,  f{a,  b)),  of  the  curve 
obtained  as  the  intersection  of  the  surface  z  =  f(x,  y)  with  the  plane  y  =  b. 
The  other  partial  derivative  ^-(a,b)  has  a  similar  geometric  interpretation.  How- 
ever, the  surface  z  =  f(x,  y)  contains  infinitely  many  curves  passing  through 
(a,  b,  f(a,  b))  whose  slope  we  might  choose  to  measure.  The  directional  deriva- 
tive enables  us  to  do  this. 


2.6  |  Directional  Derivatives  and  the  Gradient 


An  alternative  way  to  view  ^(a,b)  is  as  the  rate  of  change  of  /  as  we  move 
"infmitesimally"  from  a  =  (a,  b)  in  the  i-direction,  as  suggested  by  Figure  2.67. 
This  is  easy  to  see  since,  by  the  definition  of  the  partial  derivative, 

df<    «     r    f(a  +  h,b)-f(a,b) 
-(a,  b)  =  hm 


dx  h^O 


lim 


lim 


f((a,b)  +  (h,0))-  f(a,b) 


f((a,b)  +  h(l,0))-  f(a,b) 


o  h 

r    /(a  +  M)-/(a) 
=  lim  . 

h-+Q  h 

Note  that  we  are  identifying  the  point  (a,  b)  with  the  vector  a  =  {a,  b)  =  a\  +  bj. 
Similarly,  we  have 

V,           ..    /(a  +  /ij)-/(a) 
—  (a,  b)  =  lim  . 

dy  h^o  h 

Writing  partial  derivatives  as  we  just  have  enables  us  to  see  that  they  are 
special  cases  of  a  more  general  type  of  derivative.  Suppose  v  is  any  unit  vector  in 
R2.  (The  reason  for  taking  a  unit  vector  will  be  made  clear  later.)  The  quantity 

lim/(a  +  <,v)-/(a) 

h^O  h 

is  nothing  more  than  the  rate  of  change  of  /  as  we  move  (infmitesimally)  from  a  = 
(a ,  b)  in  the  direction  specified  by  v  =  (A,  B)  =  Ai  +  By  It's  also  the  slope  of  the 
curve  obtained  as  the  intersection  of  the  surface  z  =  f{x,y)  with  the  vertical  plane 
B(x  —  a)  —  A(y  —  b)  =  0.  (See  Figure  2.68.)  We  can  use  the  limit  expression  in 
(1)  to  define  the  derivative  of  any  scalar- valued  function  in  a  particular  direction. 


Figure  2.67  Another  way  to  view  the 
partial  derivative  9 //  dx  at  a  point. 


Figure  2.68  The  directional  derivative. 


Chapter  2  |  Differentiation  in  Several  Variables 


DEFINITION  6.1  Let  X  be  open  in  R",  /:XCR"->Ra  scalar-valued 
function,  and  a  e  X .  If  v  e  R"  is  any  unit  vector,  then  the  directional  deriva- 
tive of  /  at  a  in  the  direction  of  v,  denoted  Dv/(a),  is 

/(a  +  /*v)-/(a) 
Dv/(a)  =  hm  

h^O  h 

(provided  that  this  limit  exists). 


EXAMPLE  1  Suppose  f(x,  y)  =  x2  -  3xy  +  2x  -  5y.  Then,  if  v  =  (v,  w)  e 
R2  is  any  unit  vector,  it  follows  that 

f((0,0)  +  h(v,w))-f(0,  0) 


Dyf(0,  0)  =  lim 


h^O  h 

2v2  —  3h2vw  +  2hv  —  5hw 


=  lim 

h^o  h 

=  lim(7ii;2  —  3h  vw  +  2v  —  5w) 
=  2v  —  5w. 

Thus,  the  rate  of  change  of  /  is  2v  —  5w  if  we  move  from  the  origin  in  the 
direction  given  by  v.  The  rate  of  change  is  zero  if  v  =  (5/V29,  2/V29)  or 
(-5/V29,  -2/V59).  ♦ 

Consequently,  we  see  that  the  partial  derivatives  of  a  function  are  just  the  "tip 
of  the  iceberg."  However,  it  turns  out  that  when  /  is  differentiable,  the  partial 
derivatives  actually  determine  the  directional  derivatives  for  all  directions  v.  To 
see  this  rather  remarkable  result,  we  begin  by  defining  a  new  function  F  of  a 
single  variable  by 

F(0  =  /(a-Mv). 

Then,  by  Definition  6. 1,  we  have 

n  f(  .     Y    /(a  +  rv)-/(a)          F(t)  -  F(0) 
Dv/(a)  =  hm  =  hm  —  =  F  (0). 

t^O  t  r^O        t  —  0 

That  is, 

Ov/(a)  =  ^/(a  +  *v)|f=0.  (2) 

The  significance  of  equation  (2)  is  that,  when  /  is  differentiable  at  a,  we  can  apply 
the  chain  rule  to  the  right-hand  side.  Indeed,  let  x(r)  =  a  +  fv.  Then,  by  the  chain 
rule, 

^/(a  +  fv)  =  Df(x)Dx(t)  =  D/(x)v. 
dt 

Evaluation  at  t  =  0  gives 

Dv/(a)  =  Z)/(a)v  =  V/(a).v.  (3) 

The  purpose  of  equation  (3)  is  to  emphasize  the  geometry  of  the  situation.  The 
result  above  says  that  the  directional  derivative  is  just  the  dot  product  of  the 


2.6  |  Directional  Derivatives  and  the  Gradient 


gradient  and  the  direction  vector  v.  Since  the  gradient  is  made  up  of  the  partial 
derivatives,  we  see  that  the  more  general  notion  of  the  directional  derivative 
depends  entirely  on  just  the  direction  vector  and  the  partial  derivatives.  To  be 
more  formal,  we  summarize  this  discussion  with  a  theorem. 


THEOREM  6.2  Let  X  C  R"  be  open  and  suppose  /:  X  ->  R  is  differentiable 
at  a  e  X.  Then  the  directional  derivative  Dv/(a)  exists  for  all  directions  (unit 
vectors)  veR"  and  moreover,  we  have 

Dv/(a)  =  V/(a).v. 


EXAMPLE  2  The  function  f(x,  y)  =  x2  —  3xy  +  2x  —  5y  we  considered  in 
Example  1  has  continuous  partials  and  hence,  by  Theorem  3.5,  is  differentiable. 
Thus,  Theorem  6.2  applies  to  tell  us  that,  for  any  unit  vector  v  =  i>i  +  w\  6  R2, 

Dv/(0,  0)  =  V/(0,  0)  •  v  =  (fx(0,  0)i  +  /,(0,  0)j)  •  (vi  +  wj) 
=  (2i-5j)-(ui+Wj) 

=  2v  —  5w, 


as  seen  earlier.  ♦ 

EXAMPLE  3  The  converse  of  Theorem  6.2  does  not  hold.  That  is,  a  function 
may  have  directional  derivatives  in  all  directions  at  a  point  yet  fail  to  be  differ- 
entiable. To  see  how  this  can  happen,  consider  the  function  /:  R2  ->  R  denned 
by 


/(*.  y)  = 


if  (jc,  y)  *  (0,  0) 

xL  +  y4 

0        if(je,y)  =  (0,0) 


This  function  is  not  continuous  at  the  origin.  (Why?)  So,  by  Theorem  3.6,  it 
fails  to  be  differentiable  there;  however,  we  claim  that  all  directional  derivatives 
exist  at  the  origin.  To  see  this,  let  the  direction  vector  v  be  vi  +  wj.  Hence,  by 
Definition  6.1,  we  observe  that 


Dv/(0,  0)  =  lim 


f((0,0)  +  Kvi  +  wj))-f(0,  0) 
h 


=  lim 

h^0 


hv(hw) 


{hvf  +  (hw)4 
h2vw2 


lim 

h^o  h2(v2  +  h2w4) 

,2  „,„2 


=  lim 


vw 


h^0  V2  +  h2w4 


vw 


w 

V 


Chapter  2  |  Differentiation  in  Several  Variables 


Thus,  the  directional  derivative  exists  whenever  v  7^  0.  When  v  =  0  (in  which 
case  v  =  j),  we,  again,  must  calculate 

/((0,  0)  +  fcj)-/(0,  0) 


Dj/(0,  0)=lim 


h-*o  h 

=  hm  

fc->-0 

0-0 
=  hm  =  0. 

ft->-o  h 

Consequently,  this  directional  derivative  (which  is,  in  fact,  df/dy)  exists  as  well. 

♦ 

The  reason  we  have  restricted  the  direction  vector  v  to  be  of  unit  length  in  our 
discussion  of  directional  derivatives  has  to  do  with  the  meaning  of  Dv/(a),  not 
with  any  technicalities  pertaining  to  Definition  6.1  or  Theorem  6.2.  Indeed,  we 
can  certainly  define  the  limit  in  Definition  6. 1  for  any  vector  v,  not  just  one  of  unit 
length.  So,  suppose  w  is  an  arbitrary  nonzero  vector  in  R"  and  /  is  differentiable. 
Then  the  proof  of  Theorem  6.2  goes  through  without  change  to  give 

hm  =  V  /  (a)  •  w. 

h-*o  h 

The  problem  is  as  follows:  If  w  =  kv  for  some  (nonzero)  scalar  k,  then 

hm  =  V  /  (a)  •  w 

h-f0  h 

=  V/(a)-(*v) 
=  *(V/(a)-v) 

n  /(a  +  Av)-/(a) 


That  is,  the  "generalized  directional  derivative"  in  the  direction  of  kv  is  k  times 
the  derivative  in  the  direction  of  v.  But  v  and  kv  are  parallel  vectors,  and  it  is 
undesirable  to  have  this  sort  of  ambiguity  of  terminology.  So  we  avoid  the  trouble 
by  insisting  upon  using  unit  vectors  only  (i.e.,  by  allowing  k  to  be  ±1  only)  when 
working  with  directional  derivatives. 


Gradients  and  Steepest  Ascent   

Suppose  you  are  traveling  in  space  near  the  planet  Nilrebo  and  that  one  of  your 
spaceship's  instruments  measures  the  external  atmospheric  pressure  on  your  ship 
as  a  function  f(x,  y,  z)  of  position.  Assume,  quite  reasonably,  that  this  function 
is  differentiable.  Then  Theorem  6.2  applies  and  tells  us  that  if  you  travel  from 
point  a  =  (a ,  b,  c)  in  the  direction  of  the  (unit)  vector  u  =  ui  +  uj  +  10k,  the  rate 
of  change  of  pressure  is  given  by 

Du/(a)  =  V/(a).u. 

Now,  we  ask  the  following:  In  what  direction  is  the  pressure  increasing  the  most? 
If  9  is  the  angle  between  u  and  the  gradient  vector  V/(a),  then  we  have,  by 
Theorem  3.3  of  §1.3,  that 


Dn/(a)  =  ||V/(a)||  ||u||  costf  =  ||V/(a)||  cos#, 


2.6  |  Directional  Derivatives  and  the  Gradient  163 

since  u  is  a  unit  vector.  Because  —  1  <  cos  6  <  1 ,  we  have 

-||V/(a)||  <  Du/(a)<  ||V/(a)||. 

Moreover,  cos  9  =  1  when  9  =  0  and  cos  9  =  —  1  when  9  =  it .  Thus,  we  have 
established  the  following: 

THEOREM  6.3  The  directional  derivative  D„f(a)  is  maximized,  with  respect  to 
direction,  when  u  points  in  the  same  direction  as  V/(a)  and  is  minimized  when  u 
points  in  the  opposite  direction.  Furthermore,  the  maximum  and  minimum  values 
of  Du/(a)  are  ||  V/(a)||  and  -||  V/(a)||,  respectively. 

EXAMPLE  4   If  the  pressure  function  on  Nilrebo  is 

f(x,  y,  z)  =  5x2  +  7y4  +  x2z2  arm, 

where  the  origin  is  located  at  the  center  of  Nilrebo  and  distance  units  are  measured 
in  thousands  of  kilometers,  then  the  rate  of  change  of  pressure  at  (1,-1,2) 
in  the  direction  of  i  +  j  +  k  may  be  calculated  as  V/(l,  —  1,  2)  •  u,  where  u  = 
(i  +  j  +  k)/\/3.  (Note  that  we  normalized  the  vector  i  +  j  +  k  to  obtain  a  unit 
vector.)  Using  Theorem  6.2,  we  compute 

A./(l,-l,2)  =  V/(l,-l,2).u 

i  +  j  +  k 

=  (18i  -  28j  +  4k)  V— 

18-28  +  4 


V3 


2\/3  atm/Mm. 


Additionally,  in  view  of  Theorem  6.3,  the  pressure  will  increase  most  rapidly 
in  the  direction  of  V/(l ,  —1,2),  that  is,  in  the 

18i-28j+4k       9i-14j  +  2k 


||18i-28j+4k||  V28T 

direction.  Moreover,  the  rate  of  this  increase  is 

||  V/(l, -1,2)||  =  2^/281  atm/Mm.  ♦ 

Theorem  6.3  is  stated  in  a  manner  that  is  independent  of  dimension — that  is,  so 
that  it  applies  to  functions /:  X  C  R"  —>  R  for  any  n  >  2.  In  the  case  n  =  2,  there 
is  another  geometric  interpretation  of  Theorem  6.3:  Suppose  you  are  mountain 
climbing  on  the  surface  z  =  fix,  y).  Think  of  the  value  of  /  as  the  height  of  the 
mountain  above  (or  below)  sea  level.  If  you  are  equipped  with  a  map  and  compass 
(which  supply  information  in  the  xy-plane  only),  then  if  you  are  at  the  point  on 
the  mountain  with  xy-coordinates  (map  coordinates)  (a,  b),  Theorem  6.3  says 
that  you  should  move  in  the  direction  parallel  to  the  gradient  V/(a ,  b)  in  order  to 
climb  the  mountain  most  rapidly.  (See  Figure  2.69.)  Similarly,  you  should  move 
in  the  direction  parallel  to  —  V  f(a ,  b)  in  order  to  descend  most  rapidly.  Moreover, 
the  slope  of  your  ascent  or  descent  in  these  cases  is  ||  V/(a,  b)\\ .  Be  sure  that  you 
understand  that  V/(a,  b)  is  a  vector  in  R2  that  gives  the  optimal  north-south, 
east-west  direction  of  travel. 


1  64       Chapter  2  |  Differentiation  in  Several  Variables 


Tangent  Planes  Revisited  

In  §2.1,  we  indicated  that  not  all  surfaces  can  be  described  by  equations  of  the 
form  z  =  f(x,  y).  Indeed,  a  surface  as  simple  and  familiar  as  the  sphere  is  not  the 
graph  of  any  single  function  of  two  variables.  Yet  the  sphere  is  certainly  smooth 
enough  for  us  to  see  intuitively  that  it  must  have  a  tangent  plane  at  every  point. 
(See  Figure  2.70.) 

How  can  we  find  the  equation  of  the  tangent  plane?  In  the  case  of  the  unit 
sphere  x2  +  y2  +  z2  =  1,  we  could  proceed  as  follows:  First  decide  whether  the 
point  of  tangency  is  in  the  top  or  bottom  hemisphere.  Then  apply  equation  (4)  of 


Figure  2.70  A  sphere  and  one  of      §2.3  to  the  graph  of  z  =       —  x2  —  y2  or  z  =  —y/\  —  x2  —  y2,  as  appropriate. 

its  tangent  planes.  The  calculus  is  tedious  but  not  conceptually  difficult.  However,  the  tangent  planes 

to  points  on  the  equator  are  all  vertical  and  so  equation  (4)  of  §2.3  does  not  apply. 
(It  is  possible  to  modify  this  approach  to  accommodate  such  points,  but  we  will 
not  do  so.)  In  general,  given  a  surface  described  by  an  equation  of  the  form 
F(x,  v,  z)  =  c  (where  c  is  a  constant),  it  may  be  entirely  impractical  to  solve 
for  z  even  as  several  functions  of  x  and  y.  Try  solving  for  z  in  the  equation 
xyz  +  ye"  —  x2  +  yz2  =  0  and  you'll  see  what  we  mean.  We  need  some  other 
way  to  get  our  hands  on  tangent  planes  to  surfaces  described  as  level  sets  of 
functions  of  three  variables. 

To  get  started  on  our  quest,  we  present  the  following  result,  interesting  in  its 
own  right: 


THEOREM  6.4  Let  X  C  R"  be  open  and  /:  X  ->  R  be  a  function  of  class  C1. 
If  x0  is  a  point  on  the  level  set  S  =  {x  e  X  \  f(x)  =  c},  then  the  vector  V/(x0) 
is  perpendicular  to  S. 


2.6  |  Directional  Derivatives  and  the  Gradient 


PROOF  We  need  to  establish  the  following:  If  v  is  any  vector  tangent  to  S  at  xo, 
then  V/(xo)  is  perpendicular  to  v  (i.e.,  V/(xo)  •  v  =  0).  By  a  tangent  vector  to  S 
at  xo,  we  mean  that  v  is  the  velocity  vector  of  a  curve  C  that  lies  in  S  and  passes 
through  Xq.  The  situation  in  R3  is  pictured  in  Figure  2.71. 


Figure  2.71  The  level  set  surface 
S  =  {x  |  /(x)  =  c}. 


Thus,  let  C  be  given  parametrically  by  x(t)  =  (xi(/),  X2(t),  . . . ,  x„(t)),  where 
a  <  t  <  b  and  x(fo)  =  xo  for  some  number  to  in  (a,  b).  (Then,  if  v  is  the  velocity 
vector  at  x0,  we  must  have  x'(f0)  =  v.  See  §3.1  for  more  about  velocity  vectors.) 
Since  C  is  contained  in  S,  we  have 

/(x(0)  =  /(*i(0.  *2(0.  •  ■  • .  *»(0)  =  c- 

Hence, 

jt  imm  =  jfc\  =  o.  (4) 

On  the  other  hand  the  chain  rule  applied  to  the  composite  function  fox: 
(a,b)  ->  R  tells  us 

^[/(x(f))]  =  V/(x(0)-x'(0- 
Evaluation  at  to  and  equation  (4)  let  us  conclude  that 

V/(x(/0))-x'(*o)  =  V/(x0)-v  =  0, 
as  desired.  ■ 


Here's  how  we  can  use  the  result  of  Theorem  6.4  to  find  the  plane  tan- 
gent to  the  sphere  x2  +  y2  +  z2  =  1  at  the  point  ^—      0,  ^J.  From  §1.5,  we 

know  that  a  plane  is  determined  uniquely  from  two  pieces  of  information:  (i)  a 
point  in  the  plane  and  (ii)  a  vector  perpendicular  to  the  plane.  We  are  given  a 

point  in  the  plane  in  the  form  of  the  point  of  tangency  ^—  -7^,  0,  ^=V  As  for 

a  vector  normal  to  the  plane,  Theorem  6.4  tells  us  that  the  gradient  of  the  func- 
tion f(x,  y,  z)  =  x2  +  y2  +  z2  that  defines  the  sphere  as  a  level  set  will  do.  We 
have 

V  f(x,  y,  z)  =  2xi  +  2y\  +  2zk, 

so  that 

v/(-7!'°'7!)  =  ^'+V5k- 


Chapter  2  |  Differentiation  in  Several  Variables 


Hence,  the  equation  of  the  tangent  plane  is 

V2f,  +  -L)  +  ^--L)=0, 


=  0, 


or 


X  =  y/2. 


In  general,  if  S  is  a  surface  in  R3  defined  by  an  equation  of  the  form 

f(x,y,z)  =  c, 

then  if  xo  £  X,  the  gradient  vector  V/(xo)  is  perpendicular  to  S  and,  conse- 
quently, if  nonzero,  is  a  vector  normal  to  the  plane  tangent  to  S  at  x0.  Thus, 
the  equation 


V/(x0)-(x-x0)  =  0 

or,  equivalently, 

fx(x0,  y0,  z0)(x  -  x0)  +  fy(x0,  y0,  zo)(y  -  v0) 

+  fz(*o,  yo,  zo)(z  -  zo)  =  0 
is  an  equation  for  the  tangent  plane  to  S  at  x0. 


(5) 


(6) 


Note  that  formula  (5)  can  be  used  in  R"  as  well  as  in  R3,  in  which  case  it 
defines  the  tangent  hyperplane  to  the  hypersurface  S  C  R"  defined  by  f{x\ , 
X2,  ■  ■  ■ ,  xn)  =  c  at  the  point  xo  e  S. 

EXAMPLE  5   Considerthe  surface  S  defined  by  the  equation*3  y  —  yz2  +  z5  = 
9.  We  calculate  the  plane  tangent  to  S  at  the  point  (3,-1,2). 
To  do  this,  we  define  f(x,  y,  z)  =  x3y  —  yz2  +  z5-  Then 

V/(3,  -1,2)=  (3x2yi  +  (x3  -  z2)\  +  (5Z4  -  2yz)k)\(3  _h2) 

=  -21\  +  23j  +  84k 

is  normal  to  S  at  (3,  —1,  2)  by  Theorem  6.4.  Using  formula  (6),  we  see  that  the 
tangent  plane  has  equation 


or,  equivalently, 


-270  -  3)  +  23(y  +  1)  +  84(z  -  2)  =  0 


-27x  +  23y  +  84z  =  64. 


EXAMPLE  6  Consider  the  surface  defined  by  z4  =  x2  +  y2.  This  surface  is 
the  level  set  (at  height  0)  of  the  function 


f{x,  y,  Z)  =  x2  +  y2-  z4. 


The  gradient  of  /  is 


Vf(x,  y,z)  =  2xi  +  2yj-  4z3  k. 


2.6  |  Directional  Derivatives  and  the  Gradient 


Figure  2.72  The 

surface  of  Example  6. 


Note  that  the  point  (0,  0,  0)  lies  on  the  surface.  However,  V/(0,  0,  0)  =  0,  which 
makes  the  gradient  vector  unusable  as  a  normal  vector  to  a  tangent  plane.  Thus, 
formula  (6)  doesn't  apply.  What  we  conclude  from  this  example  is  that  the  surface 
fails  to  have  a  tangent  plane  at  the  origin,  a  fact  that  is  easy  to  believe  from  the 
graph.  (See  Figure  2.72.)  ♦ 

EXAMPLE  7   The  equation  x2  +  y2  +  z2  +  w2  =  4  defines  a  hypersphere  of 

radius  2  in  R4.  We  use  formula  (5)  to  determine  the  hyperplane  tangent  to  the 
hypersphere  at  (—  1 ,  1 ,  1,-1). 

The  hypersphere  may  be  considered  to  be  the  level  set  at  height  4  of  the 
function  f(x,  y,  z,  w)  =  x2  +  y2  +  z2  +  w2,  so  that  the  gradient  vector  is 

V f(x,  y,  z,  w)  =  (2x,  2y,  2z,  2w), 

so  that 

V/(-l,  1,1,-1)  =  (-2, 2,  2, -2). 
Using  formula  (5),  we  obtain  an  equation  for  the  tangent  hyperplane  as 
(-2,2,2,  -2)-(x  +  l,y-  1,  z  -  1,  w  +  1)  =  0 

or 

-2(x  +  1)  +  2(y  -  1)  +  2(z  -  1)  -  2(w  +  1)  =  0. 
Equivalently,  we  have  the  equation 

x  -  y  -  z  +  w  +  4  =  0.  ♦ 


EXAMPLE  8   We  determine  the  plane  tangent  to  the  paraboloid  z 


-  v2 


3y 


at  the  point  (—2,  1,  7)  in  two  ways:  (i)  by  using  formula  (4)  in  §2.3,  and  (ii)  by 
using  our  new  formula  (6). 

First,  the  equation  z  =  x2  +  3y2  explicitly  describes  the  paraboloid  as  the 
graph  of  the  function  f(x,  y)  =  x2  +  3y2,  that  is,  by  an  equation  of  the  form 
z  =  fix,  y).  Therefore,  formula  (4)  of  §2.3  applies  to  tell  us  that  the  tangent 
plane  at  (—2,  1,  7)  has  equation 

z  =  f(-2,  1)  +  fx(-2,  l)(x  +  2)  +  fy(-2,  l)(y  -  1) 

or,  equivalently, 

z  =  7  -  4(jc  +  2)  +  6(y  -  1).  (7) 

Second,  if  we  write  the  equation  of  the  paraboloid  as  x2  +  3y2  —  z  =  0, 
then  we  see  that  it  describes  the  paraboloid  as  the  level  set  of  height  0  of  the 
three-variable  function  F(x,  y,  z)  =  x2  +  3y2  —  z.  Hence,  formula  (6)  applies 
and  indicates  that  an  equation  for  the  tangent  plane  at  (—2,  1,  7)  is 

Fx(-2,  1,  7)(jc  +  2)  +  Fy{-2,  1,  7)(y  -  1)  +  Fz(-2,  1,  7)(z  -  7)  =  0 


or 


-4(x  +  2)  +  6(y-  1)  -  l(z  -  7) 
As  can  be  seen,  equation  (7)  agrees  with  equation  (8). 


(8) 

♦ 


Example  8  may  be  viewed  in  a  more  general  context.  If  S  is  the  surface  in  R3 
given  by  the  equation  z  =  f(x,  y)  (where  /  is  differentiate),  then  formula  (4)  of 
§  2 . 3  tells  us  that  an  equation  for  the  plane  tangent  to  S  at  the  point  (a ,  b ,  /  {a ,  b))  is 


z  =  f(a,  b)  +  fx(a,  b)(x  -a)+  Ua,  b)(y  -  b). 


Chapter  2  |  Differentiation  in  Several  Variables 


(-2,2,6) 


Figure  2.73  The  two-sheeted 
hyperboloid  z2/4  —  x2  —  y2  =  1. 
The  point  (—2,  2,  6)  lies  on  the 
sheet  given  by  z  =  2^/x2  +  y2  +  1, 
and  the  point  (1,1,  —  2^/3)  lies  on 
the  sheet  given  by 
z  =  -2jx2  +  y2+l. 


1. 


At  the  same  time,  the  equation  for  S  may  be  written  as 

f(x,  y)  -  z  =  0. 

Then,  if  we  let  F(x,  y,  z)  =  f(x,  y)  —  z,  we  see  that  S  is  the  level  set  of  F  at 
height  0.  Hence,  formula  (6)  tells  us  that  the  tangent  plane  at  (a,  b,  f(a,  b))  is 

Fx(a,  b,  f(a,  b))(x  -  a)  +  Fy(a,  b,  f(a,  b))(y  -  b) 

+  Fe(a,  b,  f(a,  b))(z  -  f(a,  b))  =  0. 

By  construction  of  F, 

dF  _df        dF  _  df  dF 

dx       dx '       dy       dy '  dz 

Thus,  the  tangent  plane  formula  becomes 

fx(a,  b)(x  -a)  +  fy(a,  b)(y  -  b)  -  (z  -  f(a,  b))  =  0. 

The  last  equation  for  the  tangent  plane  is  the  same  as  the  one  given  above  by 
equation  (4)  of  §2.3. 

The  result  shows  that  equations  (5)  and  (6)  extend  the  formula  (4)  of  §2.3  to 
the  more  general  setting  of  level  sets. 

The  Implicit  Function  and  Inverse  Function 

Theorems  (optional)   

We  have  previously  noted  that  not  all  surfaces  that  are  described  by  equations  of 
the  form  F(x  ,y,z)  =  c  can  be  described  by  an  equation  of  the  form  z  =  f(x ,  y). 
We  close  this  section  with  a  brief — but  theoretically  important — digression  about 
when  and  how  the  level  set  {(x,  y,  z)  |  F(x,  y,  z)  =  c}  can  also  be  described  as 
the  graph  of  a  function  of  two  variables,  that  is,  as  the  graph  of  z  =  f(x,  y). 
We  also  consider  the  more  general  question  of  when  we  can  solve  a  system  of 
equations  for  some  of  the  variables  in  terms  of  the  others. 
We  begin  with  an  example. 


EXAMPLE  9  Consider  the  hyperboloid  z2/4  -  x2 
described  as  the  level  set  (at  height  1)  of  the  function 


y  =  1,  which  may  be 


F{x,  y,  z)  = 


y 


(See  Figure  2.73.)  This  surface  cannot  be  described  as  the  graph  of  an  equation 
of  the  form  z  =  f(x,  y),  since  particular  values  for  x  and  y  give  rise  to  two  values 
for  z.  Indeed,  when  we  solve  for  z  in  terms  of  x  and  y,  we  find  that  there  are  two 
functional  solutions: 


x 2  +  y2  +  1    and  z 


=  -2/ 


x2  +  y2  +  1. 


(9) 


On  the  other  hand  these  two  solutions  show  that,  given  any  particular  point 
(*o,  yo,  zo)  of  the  hyperboloid,  we  may  solve  locally  for  z  in  terms  of  x  and  y. 
That  is,  we  may  identify  on  which  sheet  of  the  hyperboloid  the  point  {xq,  yo,  Zo) 
lies  and  then  use  the  appropriate  expression  in  (9)  to  describe  that  sheet.  ♦ 

Example  9  prompts  us  to  pose  the  following  question:  Given  a  surface  S, 
described  as  the  level  set  {(x,  y,  z)  |  F(x,  y,  z)  =  c],  can  we  always  determine  at 
least  a  portion  of  S  as  the  graph  of  a  function  z  =  f(x ,  y)?  The  result  that  follows, 


2.6  |  Directional  Derivatives  and  the  Gradient 


a  special  case  of  what  is  known  as  the  implicit  function  theorem,  provides 
relatively  mild  hypotheses  under  which  we  can. 


THEOREM  6.5  (The  implicit  function  theorem)  Let  F:XcR"->R  be 
of  class  C1  and  let  a  be  a  point  of  the  level  set  S  =  {x  6  R"  |  F(x)  =  c}.  If 
FXn(a)  0,  then  there  is  a  neighborhood  U  of  (a\,  ci2, . . . ,  a„_i)  in  Rn_1,  a 
neighborhood  V  of  an  in  R,  and  a  function  f:U<^  R"_1  ->  V  of  class  C1 
such  that  if  (xi,  X2,  ■  ■  ■ ,  xn-\)  e  U  andx„  e  V  satisfy  F(x\,  X2,  ■  •  ■ ,  xn)  =  c  (i.e., 
(x\,x2,       x„)  e  S),  thenx„  =  f(xi,x2, x„_i). 


The  significance  of  Theorem  6.5  is  that  it  tells  us  that  near  a  point  a  €  S 
such  that  dF/dxn  ^  0,  the  level  set  S  given  by  the  equation  F{x\, . . . ,  xn)  =  c 
is  locally  also  the  graph  of  a  function  x„  =  /(xj , . . . ,  x„_i).  In  other  words,  we 
may  solve  locally  for  xn  in  terms  of  x\, . . . ,  x„_i,  so  that  S  is,  at  least  locally,  a 
differentiable  hypersurface  in  R" . 

EXAMPLE  10  Returning  to  Example  9,  we  recall  that  the  hyperboloid  is  the 
level  set  (at  height  1)  of  the  function  F(x,  y,  z)  =  z2/4  —  x2  —  y2.  We  have 


dF 

9z" 


Note  that  for  any  point  (xo,  yo>  Zo)  in  the  hyperboloid,  we  have  |zol  >  2.  Hence, 
9F.(xo,  yo,  zo)  0.  Thus,  Theorem  6.5  implies  that  we  may  describe  a  portion 
of  the  hyperboloid  near  any  point  as  the  graph  of  a  function  of  two  variables.  This 
is  consistent  with  what  we  observed  in  Example  9.  ♦ 

Of  course,  there  is  nothing  special  about  solving  for  the  particular  vari- 
able x„  in  terms  of  Suppose  a  is  a  point  on  the  level  set  S  de- 
termined by  the  equation  F(x)  =  c  and  suppose  VF(a)  /  0.  Then  FXj(a)  0  for 
some  i .  Hence,  we  can  solve  locally  near  a  for  x,-  as  a  differentiable  function  of 
x\, . . . ,  x,_i,  Xi+i,  . . . ,  x„.  Therefore,  S  is  locally  a  differentiable  hypersurface 
inR". 


EXAMPLE  11  Let  S  denote  the  ellipsoid  x2/4  +  y2/36  +  z2/9 
is  the  level  set  (at  height  1)  of  the  function 


1.  Then  S 


n*,y,z)=^  +  y-  + 


At  the  point  (V2,  V6,  V3),  we  have 


dF 

~dz 


2z 


(V2.V6.V3) 


2^3 


(V2,V6,V3) 


Thus,  S  may  be  realized  near  (V2,  V6,  V3)  as  the  graph  of  an  equation  of  the 

form  z  =  f(x,  y),  namely,  z  =  3^/1  -  x2/4  —  y2/36.  At  the  point  (0,  —6,  0), 
however,  we  see  that  dF /dz  vanishes.  On  the  other  hand, 


dF 

97 


(0,-6,0) 


2± 

36 


(0,-6,0) 


^0. 


Consequently,  near  (0,  —6,  0),  the  ellipsoid  may  be  described  by  solving  for  y  as 
a  function  of  x  and  z,  namely,  y  =  —6^/ 1  —  x2/4  —  z2/9.  ♦ 


Chapter  2  |  Differentiation  in  Several  Variables 


EXAMPLE  1 2  Consider  the  set  of  points  S  defined  by  the  equation  x2z2  —  y  = 
0.  Then  S  is  the  level  set  at  height  0  of  the  function  F(x,  y,  z)  =  x2z2  —  y.  Note 
that 

VF(x,y,z)  =  (2xz2, -l,2x2z). 

Since  dF  /dy  never  vanishes,  we  see  that  we  can  always  solve  for  y  as  a  function 
of  x  and  z.  (This  is,  of  course,  obvious  from  the  equation.)  On  the  other  hand,  near 
points  where  x  and  z  are  nonzero,  both  dF/dx  and  dF/dz  are  nonzero.  Hence, 
we  can  solve  for  either  x  or  z  in  this  case.  For  example,  near  (1 ,  1 ,  —  1),  we  have 

x=H  and  z=-Ii- 

As  just  mentioned,  Theorem  6.5  is  actually  a  special  case  of  a  more  general 
result.  In  Theorem  6.5  we  are  attempting  to  solve  the  equation 

F(xi,  x2,  . . . ,  xn)  =  c 

for  xn  in  terms  of  x\, . . . ,  xn-\.  In  the  general  case,  we  have  a  system  of  m 
equations 

F\(x\,  . . . ,  x„,  yi, . . . ,  ym)  =  c1 
F2(xi,...,x„,yi,...,ym)  =  c2 

^mv^l '  •  ■  ■  '        3^1)  ■  ■  ■  '  y?n) 

and  we  desire  to  solve  the  system  for  yi , . . . ,  ym  in  terms  of  x\ , . . . ,  x„.  Using  vec- 
tor notation,  we  can  also  write  this  system  as  F(x,  y)  =  c,  where  x  =  (xi, . . . ,  xn), 
y  =  (yi, . . . ,  ym),  c  =  (ci , . . . ,  cm),  and  Fi, ... ,  Fm  make  up  the  component 
functions  of  F.  With  this  notation,  the  general  result  is  the  following: 


THEOREM  6.6  (The  implicit  function  theorem,  general  case)  Suppose 
F:  A  ->•  Rm  is  of  class  C1,  where  A  is  open  in  R"+m.  Let  (a,  b)  =  (a\, . . . ,  a„, 
b\, . . . ,  bm)  £  A  satisfy  F(a,  b)  =  c.  If  the  determinant 


r  3Fi 

dy 


A(a,  b)  =  det 


(a,b) 


9F, 

dy, 


-(a,  b) 


dFm 


L  dy 


-(a,  b) 


dFm 

dyr. 


(a,  b) 


^0, 


then  there  is  a  neighborhood  U  of  a  in  R"  and  a  unique  function  f :  U  —>  Rm  of 
class  C1  such  that  f(a)  =  b  and  F(x,  f(x))  =  c  for  all  x  e  U.  In  other  words,  we 
can  solve  locally  for  y  as  a  function  f(x). 


EXAMPLE  13  We  show  that,  near  the  point  (x\,  x%,  x^,  y\,  y2)  =  (—  1,  1,  1, 
2,  1),  we  can  solve  the  system 

x\y2  +  x2y\  =  1 
xfx3yi  +  x2y{  =  3 
for  y\  and  y2  in  terms  of  x\ ,  x2,  x%. 


2.6  |  Directional  Derivatives  and  the  Gradient 


We  apply  the  general  implicit  function  theorem  (Theorem  6.6)  to  the  system 

IFi(xi,  x2,  x3,  y\,yi)  =  x\y2  +  x2yi  =  1 
F2(xi,x2,  x3,  y\,y2)  =  xfx3yi  +  x2y\  =  3 
The  relevant  determinant  is 


A(-l,  1,  1,2,  1)  =  det 


=  det 


=  det 


3Fi  dFi 

dyi  dy2 

dF2  dF2 

dyi  dy2 

x2  X\ 


(X1,.T2,X3,3'1,}'2)=(- 1,1, 1,2,1) 


xfxj,  3x2yj 


(xi  ,a-2, x3 ,  vi , y2)={- 1 , 1 , 1 ,2, 1) 

4/0. 


Hence,  we  may  solve  locally,  at  least  in  principle. 

We  can  also  use  the  equations  in  (11)  to  determine,  for  example, 

dy2 

 (— 1,  1,  1),  where  we  treat  x\,  x2,  x3  as  independent  variables  and  y\  and 

dx\ 

y2  as  functions  of  them. 

Differentiating  the  equations  in  (1 1)  implicitly  with  respect  to  X\  and  using 
the  chain  rule,  we  obtain 


yi  +  xx- — Vx2— 

ax\  ax\ 


0 


Ixix^yi  +x\xt,-  r-3x2y; 


dy2 


=  0 


.2„ 

Now,  let  (xi ,  x2,  Xi,  y\ ,  y2)  —  (—  1,  1 ,  1,2,  1),  so  that  the  system  becomes 


f*(-l,l,l)-|^(-l,l,l)  =  -l 
3xi  6x\ 

^l(-l,l,l)  +  3^(-l,l,l)  =  4 
3xi  dxi 

3y2  5 

We  may  easily  solve  this  last  system  to  find  that  —  (—1,1,1)  =  —.  ♦ 

3xi  4 

Now,  suppose  we  have  a  system  of  n  equations  that  defines  the  variables 
yi , . . . ,  yn  in  terms  of  the  variables  xi, . . . ,  x„,  that  is, 


yi  =  /i(xi, . . .  ,x„) 
yi  =  fi(xu  ■ . .  ,xn) 

y„  =  fn(X\,...,Xn) 


(12) 


Note  that  the  system  given  in  (12)  can  be  written  in  vector  form  as  y  =  f(x).  The 
question  we  ask  is,  when  can  we  invert  this  system?  In  other  words,  when  can  we 


Chapter  2  |  Differentiation  in  Several  Variables 


solve  for  x\ ,  . . .  ,xn  in  terms  of  yi , 
function  g  so  that  x  =  g(y)? 

The  solution  is  to  apply  Theorem  6.6  to  the  system 


yn,  or,  equivalently,  when  can  we  find  a 


F2(xi, 


,xn,y\, 
x„,y{, 

.xn,y\, 


y„)  =  0 

y„)  =  o 

Jn)  =  0 


where  Ff(jti,  . . . ,  xn,  y\, . . . ,  y„)  =  fi(xi,  . . . ,  xn)  —  y,-.  (In  vector  form,  we  are 
setting  F(x,  y)  =  f(x)  —  y.)  Then  solvability  for  x  in  terms  of  y  near  x  =  a,  y  =  b 
is  governed  by  the  nonvanishing  of  the  determinant 


detDf(a)  =  det 


This  determinant  is  also  denoted  by 

3(/i- 


3xi 


dxt 


(a) 


(a) 


dxn 

dfn 
dxn 


(a) 


(a) 


d(xu  . . .  ,x„) 

and  is  called  the  Jacobian  of  f  =  (/i, . . . ,  /„).  A  more  precise  and  complete 
statement  of  what  we  are  observing  is  the  following: 


THEOREM  6.7  (The  inverse  function  theorem)    Suppose  f  =  (f\  /„) 

is  of  class  C1  on  an  open  set  A  c  R" .  If 

then  there  is  an  open  set  U  c  R"  containing  a  such  that  f  is  one-one  on  U,  the 
set  V  =  f(U)  is  also  open,  and  there  is  a  uniquely  determined  inverse  function 
g:  V  ->  U  to  f,  which  is  also  of  class  C1 .  In  other  words,  the  system  of  equations 
y  =  f(x)  may  be  solved  uniquely  as  x  =  g(y)  for  x  near  a  and  y  near  b. 


det  Df(a) 


3(/i,..../») 


3(xi,  . . . ,  xn) 


EXAMPLE  14  Consider  the  equations  that  relate  polar  and  Cartesian  coordi- 
nates: 

'  x  =  r  cos# 
y  =  r  sin  6 

These  equations  define  x  and  y  as  functions  of  r  and  6.  We  use  Theorem  6.7  to 
see  near  which  points  of  the  plane  we  can  invert  these  equations,  that  is,  solve  for 
r  and  9  in  terms  of  x  and  y. 

To  use  Theorem  6.7,  we  compute  the  Jacobian 


d(x,  y) 
d(r,  9) 


cos  6  —rsmO 
sin  0     r  cos  9 


r. 


Thus,  we  see  that,  away  from  the  origin  (r  =  0),  we  can  solve  (locally)  for  r  and  9 
uniquely  in  terms  of  x  and  y.  At  the  origin,  however,  the  inverse  function  theorem 


2.6  I  Exercises  173 


does  not  apply.  Geometrically,  this  makes  perfect  sense,  since  at  the  origin  the 
polar  angle  9  can  have  any  value.  ♦ 


2.6  Exercises 


1 .  Suppose  f(x ,  y ,  z)  is  a  differentiable  function  of  three 
variables. 

(a)  Explain  what  the  quantity  V/(x,  y,  z)  •  (— k)  rep- 
resents. 

(b)  How  does  V/(x,  y,  z)  •  (-k)  relate  to  3//3z? 

In  Exercises  2—8,  calculate  the  directional  derivative  of  the 
given  function  f  at  the  point  a  in  the  direction  parallel  to  the 
vector  u. 


2.  f(x,  y)  =  ev  sinx,  a  =        o),  u 


31- j 


10 


3.  f(x,  y)  =  x2-  2x3y  +  2y\  a  =  (2,  -1),  u  = 


4.  f{x,  y)  = 


1 


-,a  =  (3, -2),u  =  i-j 


(x2  +  y2)' 

5.  /(x,  y)  =  ex  -  x2y,  a  =  (1,  2),  u  =  2i  +  j 


6.  f(x,  y,  z)  =  xyz,  a  =  (-1,  0,  2),  u 


2k-  i 

~7f 


7.  /(x,  y,  z)  =  e~ 
8-  f(x,y,z) 


3k 

9.  For  the  function 

/(*.  y) 


3z2  +  l 


+z  \a  =  (l,2,3),u  =  i  +  j  +  k 
,  a  =  (2, -1,0),  u  =  i-2j  + 


x\y\ 


0 


if  (x,  y)  ?  (0,  0) 
if(x,y)  =  (0,  0) 


(a)  calculate  fx(0,  0)  and  fy(0,  0).  (You  will  need  to 
use  the  definition  of  the  partial  derivative.) 

(b)  use  Definition  6.1  to  determine  for  which  unit 
vectors  v  =  ui  +  urj  the  directional  derivative 
Dv/(0,  0)  exists. 

(c)  use  a  computer  to  graph  the  surface  z  =  f(x,  y). 
1 0.  For  the  function 


XV 


Jx2  +  . 


if(x,y)/(0,  0) 
if(x,y)  =  (0,  0) 


(a)  calculate  /v(0,  0)  and  /,(0,  0). 


(b)  use  Definition  6.1  to  determine  for  which  unit 
vectors  v  =  vi  +  w\  the  directional  derivative 
Dv/(0,  0)  exists. 

(c)  use  a  computer  to  graph  the  surface  z  =  f(x,  y). 

1 1 .  The  surface  of  Lake  Erehwon  can  be  represented  by  a 
region  D  in  the  jcv-plane  such  that  the  lake's  depth  (in 
meters)  at  the  point  (x,  y)  is  given  by  the  expression 
400  —  3x2y2.  If  your  calculus  instructor  is  in  the  wa- 
ter at  the  point  (1,  —2),  in  which  direction  should  she 
swim 

(a)  so  that  the  depth  increases  most  rapidly  (i.e.,  so 
that  she  is  most  likely  to  drown)? 

(b)  so  that  the  depth  remains  constant? 

12.  A  ladybug  (who  is  very  sensitive  to  temperature)  is 
crawling  on  graph  paper.  She  is  at  the  point  (3,7)  and 
notices  that  if  she  moves  in  the  i-direction,  the  tem- 
perature increases  at  a  rate  of  3  deg/cm.  If  she  moves 
in  the  j -direction,  she  finds  that  her  temperature  de- 
creases at  a  rate  of  2  deg/cm.  In  what  direction  should 
the  ladybug  move  if 

(a)  she  wants  to  warm  up  most  rapidly? 

(b)  she  wants  to  cool  off  most  rapidly? 

(c)  she  desires  her  temperature  not  to  change? 

13.  You  are  atop  Mt.  Gradient,  5000  ft  above  sea  level, 
equipped  with  the  topographic  map  shown  in  Fig- 
ure 2.74.  A  storm  suddenly  begins  to  blow,  necessitat- 
ing your  immediate  return  home.  If  you  begin  heading 
due  east  from  the  top  of  the  mountain,  sketch  the  path 
that  will  take  you  down  to  sea  level  most  rapidly. 

14.  It  is  raining  and  rainwater  is  running  off  an  ellipsoidal 
dome  with  equation  Ax2  +  y2  +  4z2  =  16,  where 
z  >  0.  Given  that  gravity  will  cause  the  raindrops  to 
slide  down  the  dome  as  rapidly  as  possible,  describe 
the  curves  whose  paths  the  raindrops  must  follow. 
(Hint:  You  will  need  to  solve  a  simple  differential 
equation.) 

15.  Igor,  the  inchworm,  is  crawling  along  graph  paper  in 
a  magnetic  field.  The  intensity  of  the  field  at  the  point 
(x,  y)  is  given  by  M(x,  y)  =  3x2  +  y2  +  5000.  If  Igor 
is  at  the  point  (8,  6),  describe  the  curve  along  which  he 
should  travel  if  he  wishes  to  reduce  the  field  intensity 
as  rapidly  as  possible. 

In  Exercises  16— 19,  find  an  equation  for  the  tangent  plane  to 
the  surface  given  by  the  equation  at  the  indicated  point  (xo ,  yo , 
zo)- 


1  74       Chapter  2  |  Differentiation  in  Several  Variables 


Figure  2.74  The  topographic  map  of  Mt.  Gradient  in  Exercise  13. 


16.  x1  +  y3  +  ;3  =  7,  (xq,  y0,  zo)  =  (0,  -1,  2) 

17.  zey  cosx  =  1,  (x0,  yo,  zo)  =  (x,  0,  -1) 

18.  2xz  +  yz-  x2y  +  10  =  0,  Oo,  yo,  zo)  =  (1,  -5,  5) 

19.  2xy2  =  2z2  -  xyz,  (x0,  y0,  Zo)  =  (2,  -3,  3) 

20.  Calculate  the  plane  tangent  to  the  surface  whose  equa- 
tion is  x2  —  2y2  +  5xz  =  7  at  the  point  (—1,  0,  —  |)in 
two  ways: 

(a)  by  solving  for  z  in  terms  of  x  and  y  and  using 
formula  (4)  in  §2.3 

(b)  by  using  formula  (6)  in  this  section. 

21.  Calculate  the  plane  tangent  to  the  surface  xsiny  + 
xz2  =  2eyz  at  the  point  (2,  j,  0)  in  two  ways: 

(a)  by  solving  for  x  in  terms  of  y  and  z  and  using  a 
variant  of  formula  (4)  in  §2.3 

(b)  by  using  formula  (6)  in  this  section. 

22.  Find  the  point  on  the  surface  x3  —  2y2  +  z2  =  21 
where  the  tangent  plane  is  perpendicular  to  the  line 
given  parametrically  as  x  =  3f  —  5,  v  =  2t  +  7,  z  = 
1  -  V2f. 

23.  Find  the  points  on  the  hyperboloid  9x2  —  45y2  + 
5z2  =  45  where  the  tangent  plane  is  parallel  to  the 
plane  x  +  5y  —  2z  =  7. 

24.  Show  that  the  surfaces  z  =  lx2  —  \2x  —  5y2  and 


xyz 


2  intersect  orthogonally  at  the  point  (2,  1 ,  —  1). 


25.  Suppose  that  two  surfaces  are  given  by  the  equations 


Moreover,  suppose  that  these  surfaces  intersect  at  the 
point  (jco,  yo,  Zo)-  Show  that  the  surfaces  are  tangent  at 
(x0,  yo,  Zo)  if  and  only  if 

VF(x0,  yo,  zo)  x  VG(i0,  y0,  zo)  =  0. 

26.  Let  S  denote  the  cone  x2  +  Ay2  =  z2  ■ 

(a)  Find  an  equation  for  the  plane  tangent  to  S  at  the 
point  (3,  —2,  —5). 

(b)  What  happens  if  you  try  to  find  an  equation  for  a 
tangent  plane  to  S  at  the  origin?  Discuss  how  your 
findings  relate  to  the  appearance  of  S. 

27.  Consider  the  surface  S  defined  by  the  equation  x3  — 

x2y2  +  Z2  =  0. 

(a)  Find  an  equation  for  the  plane  tangent  to  S  at  the 
point  (2,  -3/2,  1). 

(b)  Does  S  have  a  tangent  plane  at  the  origin?  Why  or 
why  not? 

If  a  curve  is  given  by  an  equation  of  the  form  fix,  y)  =  0,  then 
the  tangent  line  to  the  curve  at  a  given  point  (xo,  yo)  °n  it  niay 
be  found  in  two  ways:  (a)  by  using  the  technique  of  implicit 
differentiation  from  single-variable  calculus  and  (b)  by  using 
a  formula  analogous  to  formula  (6).  In  Exercises  28-30,  use 
both  of  these  methods  to  find  the  lines  tangent  to  the  given 
curves  at  the  indicated  points. 

28.  x2  +  y2  =  4,  (xo,  y0)  =  (-V2,  V2) 

29.  y3  =x2+x\(x0,yo)  =  (1,^/2) 


F(x,  y,  z)  =  c      and      G(x,  y,  z)  =  k. 


30.  x5  +  2xy  +  y3  =  16,  (x0,  y0)  =  (2,  -2) 


2.6  I  Exercises  175 


Let  C  be  a  curve  in  R2  given  by  an  equation  of  the  form 
f(x,  y)  =  0.  The  normal  line  to  C  at  a  point  (xo,  yo)  on  it 
is  the  line  that  passes  through  (xo,  yo)  and  is  perpendicular 
to  C  (meaning  that  it  is  perpendicular  to  the  tangent  line  to 
C  at  (xo,  yo)).  In  Exercises  31-33,  find  the  normal  lines  to 
the  given  curves  at  the  indicated  points.  Give  both  a  set  of 
parametric  equations  for  the  lines  and  an  equation  in  the  form 
Ax  +  By  =  C.  (Hint:  Use  gradients.) 

31.  x2-y2  =  9,(x0,y0)  =  (5,  -4) 

32.  x2  -  x3  =  y2,  (x0,  yo)  =  (-1,  72) 

33.  x3  -  2xy  +  y5  =  11,  (x0,  y0)  =  (2,  -1) 

34.  This  problem  concerns  the  surface  defined  by  the  equa- 
tion 

x3z  +  x2y2  +  sin(yz)  =  -3. 

(a)  Find  an  equation  for  the  plane  tangent  to  this  sur- 
face at  the  point  (—1,  0,  3). 

(b)  The  normal  line  to  a  surface  S  in  R3  at  a  point 
(xo,  yo,  Zo)  on  it  is  the  line  that  passes  through 
(xo,  yo,  Zo)  and  is  perpendicular  to  S.  Find  a  set 
of  parametric  equations  for  the  line  normal  to  the 
surface  given  above  at  the  point  (—1,0,  3). 

35.  Give  a  set  of  parametric  equations  for  the  normal  line  to 
the  surface  defined  by  the  equation  exy  +  exz  —  2eyz  = 
0  at  the  point  (-1,-1,-1).  (See  Exercise  34.) 

36.  Give  a  general  formula  for  parametric  equations  for 
the  normal  line  to  a  surface  given  by  the  equation 
F(x,  y,  z)  =  0  at  the  point  (xo,  yo,  zo)  on  the  surface. 
(See  Exercise  34.) 

37.  Generalizing  upon  the  techniques  of  this  section, 
find  an  equation  for  the  hyperplane  tangent  to 
the  hypersurface  sinxi  +  cosx2  +  sinx3  +  cos  X4  + 
sinxs  =  —  1  at  the  point  (jr,  jt,  3jt/2,  2jt,  2jt)  e  R5. 

38.  Find  an  equation  for  the  hyperplane  tangent  to  the 
(«  —  l)-dimensional  ellipsoid 

2,0  2,0  2  ,        ,      2     »("  +  1) 

X[  +  2x2  +  3x3  H  h  nxn  =  

at  the  point  (-1,  -1,...,  -1)  e  R". 

39.  Find  an  equation  for  the  tangent  hyperplane  to  the  (n  — 
l)-dimensional  sphere  x\  +  x\  +  ■  ■  ■  +  x2  =  1  in  R" 
at  the  point  (1  j+Jn,  \/-Jn, . . . ,  l/Vn,  —  l/*/n). 

Exercises  40—49  concern  the  implicit  function  theorems  and 
the  inverse  function  theorem  (Theorems  6.5,  6.6,  and  6. 7). 

40.  Let  S  be  described  by  z2y3  +  x2y  =  2. 

(a)  Use  the  implicit  function  theorem  to  determine 
near  which  points  S  can  be  described  locally  as 
the  graph  of  a  C1  function  z  =  /(x,  y). 


(b)  Near  which  points  can  S  be  described  (locally)  as 
the  graph  of  a  function  x  =  g(y,  z)? 

(c)  Near  which  points  can  S  be  described  (locally)  as 
the  graph  of  a  function  y  =  h(x,  z)? 

41 .  Let  S  be  the  set  of  points  described  by  the  equation 

sinxy  +  exz  +  x3y  =  1. 

(a)  Near  which  points  can  we  describe  S  as  the  graph 
of  a  C1  function  z  =  f(x,  y)?  What  is  f{x,  y)  in 
this  case? 

(b)  Describe  the  set  of  "bad"  points  of  S,  that  is,  the 
points  (xo,  yo,  zo)  €  S  where  we  cannot  describe 
S  as  the  graph  of  a  function  z  =  f(x,  y). 

(c)  Use  a  computer  to  help  give  a  complete  picture  of 
S. 

42.  Let  F(x,  y)  =  c  define  a  curve  C  in  R2.  Suppose 
(xo,  yo)  is  a  point  ofC  such  that  VF(xo,  yo)  /  0.  Show 
that  the  curve  can  be  represented  near  (xo,  yo)  as  either 
the  graph  of  a  function  y  =  f(x)  or  the  graph  of  a 
function  x  =  g(y). 

43.  Let  F(x,  y)  =  x2  —  y3,  and  consider  the  curve  C  de- 
fined by  the  equation  F(x,  y)  =  0. 

(a)  Show  that  (0,  0)  lies  on  C  and  that  Fy(0,  0)  =  0. 

(b)  Can  we  describe  C  as  the  graph  of  a  function 
y  =  /(x)?  Graph  C. 

(c)  Comment  on  the  results  of  parts  (a)  and  (b)  in  light 
of  the  implicit  function  theorem  (Theorem  6.5). 

44.  (a)  Consider  the  family  of  level  sets  of  the  function 

F(x,  y)  =  xy  +  1.  Use  the  implicit  function  theo- 
rem to  identify  which  level  sets  of  this  family  are 
actually  unions  of  smooth  curves  in  R2  (i.e.,  locally 
graphs  of  C1  functions  of  a  single  variable). 

(b)  Now  consider  the  family  of  level  sets  of 
F(x,  y,  z)  =  xyz  +  1.  Which  level  sets  of  this 
family  are  unions  of  smooth  surfaces  in  R3? 

45.  Suppose  that  F(u,  v)  is  of  class  C1  and  is  such  that 
F(-2,  1)  =  0  and  F„(-2,  1)  =  7,  F„(-2,  1)  =  5.  Let 
G(x,  y,  z)  =  F(x3  -  2y2  +  z5,  xy  -  x2z  +  3). 

(a)  Check  that  G(-l,  1,  1)  =  0. 

(b)  Show  that  we  can  solve  the  equation  G(x,  y,  z)  = 
0  for  z  in  terms  of  x  and  y  (i.e.,  as  z  =  g(x,  y),  for 
(x,y)near(-l,  1)  so  that  g(-l,  1)=  1). 

46.  Can  you  solve 

X2j2  —  x\  cos  yi  =  5 
X2  sin  yi  +  x\yi  =  2 

for  yi,  y2  as  functions  of  x\,  xi  near  the  point 
(x\,  X2,  yu  y2)  =  (2, 3,  n,  1)?  What  about  near  the 
point  (xi ,  X2,  yi ,  ya)  =  (0,  2,  n/2,  5/2)? 


Chapter  2  |  Differentiation  in  Several  Variables 


x\y\ 


47.  Consider  the  system 

2x2y3  =  1 
xiy{  +  x2y2  -  4y2yi  =  -9  . 
x2y\  +  3xiyj  =  12 

(a)  Show  that,  near  the  point  (x\,  x%,  y\,  y%,  y{)  = 
(1,  0,  —1,  1,  2),  it  is  possible  to  solve  for  yi,  yj, 
ys  in  terms  of  x\,  x%. 

(b)  From  the  result  of  part  (a),  we  may  consider  y\,y%, 
j3  to  be  functions  of.ti  and.*^.  Use  implicit  differ- 

dy\ 

entiation  and  the  chain  rule  to  evaluate  (1,  0), 

dx\ 

-i(l,0),and-^(l,0). 
9xi  dx\ 

48.  Consider  the  equations  that  relate  cylindrical  and 
Cartesian  coordinates  in  R3: 

x  =  r  cos  8 
y  =  r  sin  8 . 

z  =  z 


(a)  Near  which  points  of  R3  can  we  solve  for  r,  8,  and 
z  in  terms  of  the  Cartesian  coordinates? 

(b)  Explain  the  geometry  behind  your  answer  in 
part  (a). 

49.  Recall  that  the  equations  relating  spherical  and  Carte- 
sian coordinates  in  R3  are 

'  x  =  p  sin  <p  cos  9 

y  =  p  sinip  sin#. 

Z  =  p  cos  <p 

(a)  Near  which  points  of  R3  can  we  solve  for  p,  <p,  and 
8  in  terms  of  x,  y,  and  z? 

(b)  Describe  the  geometry  behind  your  answer  in 
part  (a). 


Figure  2.75  The  tangent  line  to 
y  =  f(x)  at  (xq,  f(xo))  crosses  the 
x-axis  at  x  =  x\. 


2.7   Newton's  Method  (optional) 

When  you  studied  single-variable  calculus,  you  may  have  learned  a  method,  known 
as  Newton's  method  (or  the  Newton-Raphson  method),  for  approximating  the 
solution  to  an  equation  of  the  form  f(x)  =  0,  where  /:  X  C  R  —>  R  is  a  differ- 
entiable  function.  Here's  a  reminder  of  how  the  method  works. 

We  wish  to  find  a  number  r  such  that  f(r  )  =  0.  To  approximate  r,  we  make 
an  initial  guess  xq  for  r  and,  in  general,  we  expect  to  find  that  f(xo)  ^  0.  So  next 
we  look  at  the  tangent  line  to  the  graph  of  /  at  (x0,  f(x0)).  (See  Figure  2.75.) 
Since  the  tangent  line  approximates  the  graph  of  /  near  (xo,  f(xo)),  we  can 
find  where  the  tangent  line  crosses  the  x-axis.  The  crossing  point  (jci,  0)  will 
generally  be  closer  to  (r,  0)  than  (xq,  0)  is,  so  we  take  x\  as  a  revised  and  improved 
approximation  to  the  root  r  of  f{x)  =  0. 

To  find  x\ ,  we  begin  with  the  equation  of  the  tangent  line 

y  =  f(*o)  +  /'C*o)(*  -  *o), 
then  set  y  =  0  to  find  where  this  line  crosses  the  x  -axis.  Thus,  we  solve  the  equation 

/(*o)  +  f'(x0)(xi  -  x0)  =  0 

for  x\  to  find  that 

f(xo) 

X]  —  Xr>  . 

f'(x0) 

Once  we  have  x\,  we  can  start  the  process  again  using  x\  in  place  of  x0  and 
produce  what  we  hope  will  be  an  even  better  approximation  x^  via  the  formula 

fix,) 

Xi  =  X\  . 

f'(xi) 

Indeed,  we  may  iterate  this  process  and  define  Xk  recursively  by 

f{Xk-\) 


xt  =  Xk-i  k  =  1,2,... 

and  thereby  produce  a  sequence  of  numbers  xq,  x\,  . . . ,  Xk, 


(1) 


2.7  |  Newton's  Method  (optional)  177 


It  is  not  always  the  case  that  the  sequence  {xk\  converges.  However,  when 
it  does,  it  must  converge  to  a  root  of  the  equation  f(x)  =  0.  To  see  this,  let 
L  =  lim^oo  Xk.  Then  we  also  have  lim^oo  jc&_i  =  L.  Taking  limits  in  formula 
(1),  we  find 

L=L-m, 

/'(!)' 

which  immediately  implies  that  f(L)  =  0.  Hence,  L  is  a  root  of  the  equation. 

Now  that  we  have  some  understanding  of  derivatives  in  the  multivariable 
case,  we  turn  to  the  generalization  of  Newton's  method  for  solving  systems  of  n 
equations  in  n  unknowns.  We  may  write  such  a  system  as 

fi(xi,  ...,x»)  =  0 

fi(x\, x„)  =  0 

.  ■  (2) 

f„(x\,  ...,*„)  =  0 

We  consider  the  map  f:  X  c  R"  ->  R"  defined  as  f(x)  =  (/i(x), . . . ,  /n(x))  (i.e., 
f  is  the  map  whose  component  functions  come  from  the  equations  in  (2).  The 
domain  X  of  f  may  be  taken  to  be  the  set  where  all  the  component  functions  are 
denned.)  Then  to  solve  system  (2)  means  to  find  a  vector  r  =  (rj , . . . ,  r„)  such 
that  f(r)  =  0.  To  approximate  such  a  vector  r,  we  may,  as  in  the  single- variable 
case,  make  an  initial  guess  xo  for  what  r  might  be.  If  f  is  differentiable,  then  we 
know  that  y  =  f(x)  is  approximated  by  the  equation 

y  =  f(x<,)  +  Of(x0)(x  -  xo). 

(Here  we  think  of  f(xo)  and  the  vectors  x  and  xo  as  n  x  1  matrices.)  Then  we  set 
y  equal  to  0  to  find  where  this  approximating  function  is  zero.  Thus,  we  solve  the 
matrix  equation 

f(x„)  +  £>f(x0)(x1  -  x0)  =  0  (3) 

for  xi  to  give  a  revised  approximation  to  the  root  r.  Evidently  (3)  is  equivalent  to 

Df(xo)(Xl  -  x0)  =  -f(x0).  (4) 

To  continue  our  argument,  suppose  that  Df(xo)  is  an  invertible  n  x  n  matrix, 
meaning  that  there  is  a  second  n  x  n  matrix  [Df(xo)]_1  with  the  property  that 
[Df^r'Dftxo)  =  JDf(x0)[Df(x0)]"1  =  /„,  the  n  x  n  identity  matrix.  (See  Ex- 
ercises 20  and  30-38  in  §1.6.)  Then  we  may  multiply  equation  (4)  on  the  left  by 
[Df(x0)]_1  to  obtain 

7„(X!  -x0)=-[Df(x0)r1f(x0). 

Since  I„A  =  A  for  any  n  x  k  matrix  A,  this  last  equation  implies  that 

Xl  =x0-[£>f(x0)r1f(x0).  (5) 

As  we  did  in  the  one -variable  case  of  Newton's  method  we  may  iterate  formula 
(5)  to  define  recursively  a  sequence  {x^}  of  vectors  by 


xt  =  xn-[Df(Xn)r'%-i)  (6) 


Chapter  2  |  Differentiation  in  Several  Variables 


-3  + 

Figure  2.76  Finding 
the  intersection 
points  of  the  circle 


4  and 


the  hyperbola 
Ax2  -  y2  =  A  in 
Example  1. 


Note  the  similarity  between  formulas  (1)  and  (6).  Moreover,  just  as  in  the  case 
of  formula  (1),  although  the  sequence  {xo,  Xi, . . . ,  xk, . . .}  may  not  converge,  if 
it  does,  it  must  converge  to  a  root  of  f(x)  =  0.  (See  Exercise  4.) 

EXAMPLE  1  Consider  the  problem  of  finding  the  intersection  points  of  the  cir- 
cle x2  +  y2  =  4  and  the  hyperbola  4x2  —  y2  =  4.  (See  Figure  2.76.)  Analytically, 
we  seek  simultaneous  solutions  to  the  two  equations 

x2  +  y2  =  4     and     Ax2  -  y2  =  4, 

or,  equivalently,  solutions  to  the  system 


x2  +  y2  -  4  =  0 


Ax2  -  y2 


0 


(7) 


To  use  Newton's  method,  we  define  a  function  f :  R  ->  R  by  f(x ,  y)  =  (x  + 


y 


A  Ax1 


A)  and  try  to  approximate  solutions  to  the  vector  equation 


f(x,  y)  =  (0,  0).  We  may  begin  with  any  initial  guess,  say, 


x0 


Xo 

T 

l 

and  then  produce  successive  approximations  Xi,  x2, 
mula  (6).  In  particular,  we  have 


to  a  solution  using  for- 


NotethatdetDf(x,  y) 


2x 
8.v 


2y 
-2y 


Df(x,y)  = 

-20xy .  You  may  verify  (see  Exercise  36  in  §  1 .6)  that 


[Df(x,y)Tl  = 


1 


-20xy 


-2y  -2y 
-8x  2x 


1 

2 

57 


1 
1 


Thus, 


~xk~ 

~Xk-\ 

yk-\_ 

~xk-\ 

yk-i 

[Df(xk_ uyk-i )]  f(*k- 1 .  yk-i) 
l  l 


2 


L  5yk_ 


1 

10%-i  J 


4-i  +  yLi-4 

^U-yU-A 


Jxk-\ 


L     I0yk-i  J 


Xk-l 

yk-\  - 


JXk-\ 


lOxjt  i 
10W-i  J 


2.7  |  Newton's  Method  (optional)  179 


Beginning  with  xo  =  yo  =  1 ,  we  have 

5  ■  l2  —  8  5  •  l2  —  12 

xx  =  \  =  1.3  yi  =  1  =  1.7 

10-1  10-1 

5(1.3)2-8  5(1.7)2  -  12 

x2  =  1.3   =  1.265385    y2  =  1.7  

10(1.3)  y  10(1.7) 

=  1.555882,  etc. 

It  is  also  easy  to  hand  off  the  details  of  the  computation  to  a  calculator  or  a 
computer.  One  finds  the  following  results: 


k 

xk 

y* 

0 

1 

l 

1 

1.3 

1.7 

2 

1.26538462 

1.55588235 

3 

1.26491115 

1.54920772 

4 

1.26491106 

1.54919334 

5 

1.26491106 

1.54919334 

Thus,  it  appears  that,  to  eight  decimal  places,  an  intersection  point  of  the  curves 
is  (1.26491106,  1.54919334). 

In  this  particular  example,  it  is  not  difficult  to  find  the  solutions  to  (7)  exactly. 
We  add  the  two  equations  in  (7)  to  obtain 


5xz  -  8  =  0 


x2  = 


Thus,  x  =  ±V8/5.  If  we  substitute  these  values  for  x  into  the  first  equation  of 
(7),  we  obtain 


I  +  r  ~  4  =  0  y 

Hence,  y  =  ±y/l2/5.  Therefore,  the  four  intersection  points  are 


11 

5  ■ 


Since  78/5  «  1.264911064  and  ^/T2/5  «  1.54919334,  we  see  that  Newton's 
method  provided  us  with  an  accurate  approximate  solution  very  quickly.  ♦ 

EXAMPLE  2   We  use  Newton's  method  to  find  solutions  to  the  system 


x3  -  5x2  +  2x  -  y  +  13  =  0 
x3  +x2  -  14x  -  y  -  19  =  0 


(8) 


As  in  the  previous  example,  we  define  f :  R2 
y  +  13,  x3  +  x2  -  14x  -  y  -  19).  Then 


R2byf(x,  y)  =  (x3  -  5x2  +  2x 


Df(x,y)  = 


3x2  -  lOx  +  2 
3x2  +  2x  -  14 


1  80       Chapter  2  |  Differentiation  in  Several  Variables 


so  that  det  Di(x,  y)  =  \2x  —  16  and 


1 


1 


[Df(x,y)Y 


Thus,  formula  (6)  becomes 


1 


I2x  -  16 
-3x2  -  2x  +  14 
12x  -  16 


xk 

~xk-\ 

\2x  -  16 
-3x2  -lOx  +  2 
\2x  -  16 


1 


12xt. 


-3x 


16 

2*fc_i  +  14 


-3x 


12xt. 

2 


i  -  16 

lOjCjt-i  +2 


12x^-1  —  16 
5x?_,  +  2xk-\  - 


12xt_ 


<r-l 


16 


+  X 


k-1 


I4xk- 


y*-i  + 13 
y*-i  -  19 


6r2 


I6xk-i  —  32 


3.v 


16.T 


\Ax\_x 


\2xk-\  —  16 

82xt_i  -  8y*_i  +  6jCjt_iyjt_i  +  72 


6^yt-l  —  8 


This  is  the  formula  we  iterate  to  obtain  approximate  solutions  to  (8). 

If  we  begin  with  x0  =  (x0,  y0)  =  (8,  10),  then  the  successive  approximations 
\k  quickly  converge  to  (4,  5),  as  demonstrated  in  the  table  below. 


k 

Xk 

yk 

0 

8 

10 

1 

5.2 

-98.2 

2 

4.1862069 

-2.7412414 

3 

4.00607686 

4.82161865 

4 

4.00000691 

4.99981073 

5 

4.00000000 

5.00000000 

6 

4.00000000 

5.00000000 

If  we  begin  instead  with  x0  =  (50,  60),  then  convergence  is,  as  you  might  predict, 
somewhat  slower  (although  still  quite  rapid): 


k 

Xk 

yk 

0 

50 

60 

1 

25.739726 

-57257.438 

2 

13.682211 

-7080.8238 

3 

7.79569757 

-846.58548 

4 

5.11470969 

-86.660453 

5 

4.1643023 

-1.6486813 

6 

4.00476785 

4.86119425 

7 

4.00000425 

4.99988349 

8 

4.00000000 

5.00000000 

9 

4.00000000 

5.00000000 

2.7  I  Exercises  181 


On  the  other  hand,  if  we  begin  with  xo  =  (—2,  12),  then  the  sequence  of  points 
generated  converges  to  a  different  solution,  namely,  (—4/3,  —25/27): 


k 

Xk 

yk 

0 

-2 

12 

1 

-1.4 

1.4 

2 

-1.3341463 

-0.903122 

3 

-1.3333335 

-0.9259225 

4 

-1.3333333 

-0.9259259 

5 

-1.3333333 

-0.9259259 

In  fact,  when  a  system  of  equations  has  multiple  solutions,  it  is  not  always 
easy  to  predict  to  which  solution  a  given  starting  vector  xo  will  converge  under 
Newton's  method  (if,  indeed,  there  is  convergence  at  all).  ♦ 

Finally,  we  make  two  remarks.  First,  if  at  any  stage  of  the  iteration  process  the 
matrix  Df(xk)  fails  to  be  invertible  (i.e.,  [Df(x,t)]_1  does  not  exist),  then  formula 
(6)  cannot  be  used.  One  way  to  salvage  the  situation  is  to  make  a  different  choice 
of  initial  vector  xo  in  the  hope  that  the  sequence  {x^}  that  it  generates  will  not 
involve  any  noninvertible  matrices.  Second,  we  note  that  if,  at  any  stage,  x,t  is 
exactly  a  root  of  f(x)  =  0,  then  formula  (6)  will  not  change  it.  (See  Exercise  7). 


2.7  Exercises 


1 .  Use  Newton's  method  with  initial  vector  xo  =  (1 ,  —  1) 
to  approximate  the  real  solution  to  the  system 

y  V  =  3 

2yex  +  10y4  =  0  ' 

2.  In  this  problem,  you  will  use  Newton's  method  to 
estimate  the  locations  of  the  points  of  intersection 
of  the  ellipses  having  equations  3x2  +  y2  =  7  and 
x2  +4y2  =  8. 

(a)  Graph  the  ellipses  and  use  your  graph  to  give  a  very 
rough  estimate  (xo,  yo)  of  the  point  of  intersection 
that  lies  in  the  first  quadrant. 

(b)  Denote  the  exact  point  of  intersection  in  the  first 
quadrant  by  (X,  Y).  Without  solving,  argue  that 
the  other  points  of  intersection  must  be  (—X,  Y), 
(X, -Y),  and  (—X,  —Y). 

(c)  Now  use  Newton's  method  with  your  estimate 
(xq,  yo)  in  part  (a)  to  approximate  the  first  quadrant 
intersection  point  (X,  Y). 

(d)  Solve  for  the  intersection  points  exactly,  and  com- 
pare your  answer  with  your  approximations. 

3.  This  problem  concerns  the  determination  of  the  points 
of  intersection  of  the  two  curves  with  equations  jc3  — 
4y3  =  1  and  x2  +  4y2  =  2. 


(a)  Graph  the  curves  and  use  your  graph  to  give  rough 
estimates  for  the  points  of  intersection. 

(b)  Now  use  Newton's  method  with  different  initial 
estimates  to  approximate  the  intersection  points. 

4.  Consider  the  sequence  of  vectors  xo,  Xi,  . . . ,  where, 
for  k  >  1,  the  vector  x*  is  defined  by  the  Newton's 
method  recursion  formula  (6)  given  an  initial  "guess" 
xo  at  a  root  of  the  equation  f(x)  =  0.  (Here  we  as- 
sume that  f:XC  R"  — >  R"  is  a  differentiable  func- 
tion.) By  imitating  the  argument  in  the  single-variable 
case,  show  that  if  the  sequence  {xk}  converges  to  a  vec- 
tor L  and  Z)f(L)  is  an  invertible  matrix,  then  L  must 
satisfy  f(L)  =  0. 

5.  This  problem  concerns  the  Newton's  method  iteration 
in  Example  1 . 

(a)  Use  initial  vector  xo  =  (—  1,  1)  and  calculate  the 
successive  approximations  Xi,  X2,  X3,  etc.  To  what 
solution  of  the  system  of  equations  (7)  do  the  ap- 
proximations converge? 

(b)  Repeat  part  (a)  with  xo  =  (1,  —  1).  Repeat  again 
with  xo  =  (—  1,  —1). 

(c)  Comment  on  the  results  of  parts  (a)  and  (b)  and 
whether  you  might  have  predicted  them.  Describe 
the  results  in  terms  of  Figure  2.76. 


1  82       Chapter  2  |  Differentiation  in  Several  Variables 


8.  Suppose  that  f:XCR2->R2  is  differentiable  and 
that  we  write  f(x,  y)  =  (f(x,  y),  g(x,  y)).  Show  that 
formula  (6)  implies  that,  for  k  >  1, 

f(xk-i,  yk-i)gy(xk-U  Vk-l)  ~  g(Xk-l,  yk-\)fy(xk-\,  yk-i) 
fx(xk-i,  yk-i)gy{xk-i,  jh)  -  fy(xk-u  yk-i)gx(xk-u  yk-\) 

g(xk_i,yk_\)fx{xk-\,  yk-i)  -  /fa-i,  yk-i)gx(xk-U  yk-i) 
fx(xk-u  yk-i)gy(xk-u  yk-i)  -  fy(xk-u  yk-\)gx(xk-u  yk-i) 


6.  Consider  the  Newton's  method  iteration  in  Example  2. 

(a)  Use  initial  vector  xo  =  (1.4,  10)  and  calculate  the 
successive  approximations  xi,  X2,  X3,  etc.  To  what 
solution  of  the  system  of  equations  (8)  do  the  ap- 
proximations converge? 

^  (b)  Repeat  part  (a)  with  x0  =  (1.3,  10). 

(c)  In  Example  2  we  saw  that  (4,  5)  was  a  solution  of 
the  given  system  of  equations.  Is  (1.3,  10)closerto 
(4,  5)  or  to  the  limiting  point  of  the  sequence  you 
calculated  in  part  (b)? 

(d)  Comment  on  your  observations  in  part  (c).  What 
do  these  observations  suggest  about  how  easily  you 
can  use  the  initial  vector  xo  to  predict  the  value  of 
lim^oo  xk  (assuming  that  the  limit  exists)? 

7.  Suppose  that  at  some  stage  in  the  Newton's  method  it- 
eration using  formula  (6),  we  obtain  a  vector  xk  that  is 
an  exact  solution  to  the  system  of  equations  (2).  Show 
that  all  the  subsequent  vectors  x^+i,  Xk+2,  ■  •  ■  are  equal 
to  X£.  Hence,  if  we  happen  to  obtain  an  exact  root  via 
Newton's  method,  we  will  retain  it. 


9.  As  we  will  see  in  Chapter  4,  when  looking  for  maxima 
and  minima  of  a  differentiable  function  F:XCR"^ 
R,  we  need  to  find  the  points  where  DF{x\ , . . . ,  xn)  = 
[0  •  •  •  0],  called  critical  points  of  F.  Let  F(x,  y)  = 
4  sin  (xy)  +  x3  +  y 3 .  Use  Newton's  method  to  approx- 
imate the  critical  point  that  lies  near  (x ,  y)  =  (—1,  —1). 

1 0.  Consider  the  problem  of  finding  the  intersection  points 
of  the  sphere  x2  +  y2  +  7}  =  4,  the  circular  cylinder 
x2  +  y2  =  1,  and  the  elliptical  cylinder  4y2  +  z2  =  4. 

(a)  Use  Newton's  method  to  find  one  of  the  intersec- 
tion points.  By  choosing  a  different  initial  vector 
xo  =  (xq,  yo,  zo),  approximate  a  second  intersec- 
tion point.  (Note:  You  may  wish  to  use  a  computer 
algebra  system  to  determine  appropriate  inverse 
matrices.) 

(b)  Find  all  the  intersection  points  exactly  by  means  of 
algebra  and  compare  with  your  results  in  part  (a). 


True/False  Exercises  for  Chapter  2 


1 .  The  component  functions  of  a  vector- valued  function  8. 
are  vectors. 


2.  The  domain  of  f(x,  y)  =  |^x  +  y  +  1 
{(x,y)GR2  |  y  /0,x  #  v}. 


3  x 


3.  The  range  of  f(x,  y)  =  yx2  +  yL  +  1 
{(u,  v,  w)  G  R3  |  u  >  1}. 


x  +  y  y 


3  x 


x  +  y  y 


4.  The  function  f:  R3  -  {(0,  0,  0)}  ->  R3,  f(x)  =  2x/||x|| 
is  one-one. 

5.  The  graph  of  x  =  9y2  +  z2/4  is  a  paraboloid. 

6.  The  graph  of  z  +  x2  =  y2  is  a  hyperboloid. 


The  graph  of  any  function  of  two  variables  is  a  level  set  of 
a  function  of  three  variables. 

The  level  set  of  any  function  of  three  variables  is  the  graph 
of  a  function  of  two  variables. 


10. 


lim 


2v2 


(x,v)^(0,0)  x2  +  y2 


1. 


11.  If/(x,y) 


4  4 

y  —  x 
x2  +  y2 

2  when  (x,y)  =  (0,0) 


when(x,y)/(0,0\  then  /  is 


continuous. 


12. 


7.  The  level  set  of  a  function  f(x,y,z)  is  either  empty  13. 
or  a  surface. 


If  f(x,y)  approaches  a  number  L  as  (x,  y)  — >  (a,  b) 
along  all  lines  through  (a,b),  then  lim(x  y)^(aj,)f(x,y) 
=  L. 

If  limx^a  f(x)  exists  and  is  finite,  then  f  is  continuous 
at  a. 


Miscellaneous  Exercises  for  Chapter  2 


14.  fx(a,b)=  lim 


/(x,  5)  -  /(a,  fc) 


15.  If  f(x,  y,  z)  =  siny,  then  V/(x,  y,  z)  =  cosy. 

16.  If  f:R3  -*  R4  is  differentiable,  then  Df(x)  is  a  3  x  4 
matrix. 

17.  Iff  is  differentiable  at  a,  then  f  is  continuous  at  a. 

18.  If  f  is  continuous  at  a,  then  f  is  differentiable  at  a. 

1 9.  If  all  partial  derivatives  3  f/dx\ ,  . . . ,  3//  3x„  of  a  func- 
tion f(xi , . . . ,  xn)  exist  at  a  =  (ai , . . . ,  a„),  then  /  is 
differentiable  at  a. 

20.  Iff:  R4  -+  R5  and  g:  R4  -+  R5  are  both  differentiable 
at  a  G  R4,  then  D(f-  g)(a)  =  Df(a)  -  Dg(a). 

21.  There's   a   function   /   of  class   C2   such  that 

3/       i  3/  7 

—  =  y  —  2x  and  —  =  y  —  3xy  . 
ox  dy 

22.  If  the  second-order  partial  derivatives  of  /  exist  at 
(a,  b),  then  fxy(a,  b)  =  fyx(a,  b). 

23.  If  w  =  F(x,  y,  z)  and  z  =  g(x,  y)  where  F  and  g  are 
differentiable,  then 

dw  dF  dF  dg 
dx       dx       dz  dx 


24.  The  tangent  plane  to  z  =  x1  /(y  +  1)  at  the  point 
(—2,  0,  —8)  has  equation  z  =  12x  +  8y  +  16. 

25.  The  plane  tangent  to  xy/z2  =  1  at  (2,  8,  —4)  has  equa- 
tion 4x  +  y  +  2z  =  8. 

26.  The  plane  tangent  to  the  surface  x2  +  xyez  +  y3  = 
1  at  the  point  (2,-1,0)  is  parallel  to  the  vector 
3i+5j-3k. 

3/ 


27.  Dj/(x,y,Z): 


28.  D.kf(x,  y,  z) 


dy 


3/ 
dz' 


29.  If  f(x,  y)  =  sinx  cosy  and  v  is  a  unit  vector  in  R2, 
thenO^Dv/^-,-)  <  — . 

30.  If  v  is  a  unit  vector  in  R3  and  f(x,  y,  z)  =  sinx  — 
cosy  +  sinz,  then 

-V3  <  Dyf(x,y,z)  <  <&. 


Miscellaneous  Exercises  for  Chapter  2 


1.  Letf(x)  =  (i  +  k)  x  x. 

(a)  Write  the  component  functions  of  f. 

(b)  Describe  the  domain  and  range  of  f. 


order.  Complete  the  following  table  by  matching  each 
function  in  the  table  with  its  graph  and  plot  of  its  level 
curves. 


Graph 

Level  curves 

2.  Let  f(x)  =  proj3i_2j+kx,  where  x  =  xi  +  yj  +  zk. 

Function 

(uppercase 

(lowercase 

(a)  Describe  the  domain  and  range  of  f. 

fix,  y) 

letter) 

letter) 

(b)  Write  the  component  functions  of  f. 

f(*,y)  = 

1 

x2  +  y2  +  1 

f(x,  y)  = 

sin  y/x2  +  y2 

3.  Let  f(x,  y)  =  ^/xy. 

fix,  y)  = 

(3y2  -  2x2)e--r2-2>'2 

(a)  Find  the  domain  and  range  of  /. 

fix,  y)  = 

3       1  2 

y  —  3x  y 

(b)  Is  the  domain  of  /  open  or  closed?  Why? 

fix,  y)  = 

x2y2e-x*-2? 

fix,  y)  = 

-x2-v2 

ye  y 

4.  Letg(x,  y)  = 

(a)  Determine  the  domain  and  range  of  g. 

(b)  Is  the  domain  of  g  open  or  closed?  Why? 

5.  Figure  2.77  shows  the  graphs  of  six  functions  /(x,  y) 
and  plots  of  the  collections  of  their  level  curves  in  some 


6.  Consider  the  function  /(x,  y)  =  2  +  ln(x2  +  y2). 

(a)  Sketch  some  level  curves  of  /.  Give  at  least  those 
at  heights,  0,  1,  and  2.  (It  will  probably  help  if  you 
give  a  few  more.) 

(b)  Using  part  (a)  or  otherwise,  give  a  rough  sketch  of 
the  graph  of  z  =  f(x,  y). 


1  84       Chapter  2  |  Differentiation  in  Several  Variables 


J  

 1 

1  

-2-10      1  2 
Figure  2.77  Figures  for  Exercise  5. 


7.  Use  polar  coordinates  to  evaluate 

,2 


lim 


yx  -y 


(x,y)^{0,0)  x2  +  y2 

8.  This  problem  concerns  the  function 
i  2xy 

f(x,y)-- 

if(x,y) 


C2  +  V2 

0 


if(x,y)  ^(0,0) 
(0,0) 


(a)  Use  polar  coordinates  to  describe  this  function. 

(b)  Using  the  polar  coordinate  description  obtained  in 
part  (a),  give  some  level  curves  for  this  function. 

(c)  Prepare  a  rough  sketch  of  the  graph  of  /. 

(d)  Determine  lim(A:  ,,)^(o,o)  f(x ,  y),  if  it  exists. 

(e)  Is  /  continuous?  Why  or  why  not? 


9.  Let 


xy(xy  +  x2) 


F(x,  y) 


if(x,y)#(0,0) 


x4  +  y4 

0  if  (x,  y)  =  (0, 0) 


10. 


Show  that  the  function  g(x )  =  F (x ,  0)  is  continuous  at 
x  =  0.  Show  that  the  function  h(y)  =  F(0,  y)  is  con- 
tinuous at  y  =  0.  However,  show  that  F  fails  to  be 
continuous  at  (0,  0).  (Thus,  continuity  in  each  variable 
separately  does  not  necessarily  imply  continuity  of  the 
function.) 

Suppose  /:  U  c  R"  R  is  not  defined  at  a  point 
a  G  R"  but  is  defined  for  all  x  near  a.  In  other  words, 
the  domain  U  of  /  includes,  for  some  r  >  0,  the  set 
Br  =  {x  G  R"  |  0  <  || x  -  a||  <  r\.  (The  set  Br  is  just 
an  open  ball  of  radius  r  centered  at  a  with  the  point 


Miscellaneous  Exercises  for  Chapter  2 


a  deleted.)  Then  we  say  limx^a  /W  =  +00  if  f(x) 
grows  without  bound  as  x  ->  a.  More  precisely,  this 
means  that  given  any  N  >  0  (no  matter  how  large), 
there  is  some  S  >  0  such  that  if  0  <  ||x  —  a||  <  S  (i.e., 
ifx  €  fi,-),  then /(a)  >  N. 

(a)  Using  intuitive  arguments  or  the  preceding  tech- 
nical definition,  explain  why  lim^o  l/x2  =  00. 

(b)  Explain  why 

2 

lim   r   =  00. 

(x,y)^(l,3)  (x  -  l)2  +  (y  -  3)2 

(c)  Formulate  a  definition  of  what  it  means  to  say  that 

lim  /(x)  =  —00. 

x— *a 

(d)  Explain  why 

1  -x 

lim     —  ;          =  —00. 

(x,j.)^(0,0)  xy4  -  y4  +  x3  -  x2 

Exercises  11—1 7  involve  the  notion  of  windchill  temperature — 
see  Example  7  in  §2. 1,  and  refer  to  the  table  of  windchill  values 
on  page  85. 

11.  (a)  Find  the  windchill  temperature  when  the  air  tem- 

perature is  25  °F  and  the  windspeed  is  10  mph. 

(b)  If  the  windspeed  is  20  mph,  what  air  temperature 
causes  a  windchill  temperature  of  —  1 5  °F? 

12.  (a)  If  the  air  temperature  is  1 0  °F,  estimate  (to  the  near- 

est unit)  what  windspeed  would  give  a  windchill 
temperature  of  —5  °F. 

(b)  Do  you  think  your  estimate  in  part  (a)  is  high  or 
low?  Why? 

13.  At  a  windspeed  of  30  mph  and  air  temperature  of  35  °F, 
estimate  the  rate  of  change  of  the  windchill  tempera- 
ture with  respect  to  air  temperature  if  the  windspeed  is 
held  constant. 

14.  At  a  windspeed  of  1 5  mph  and  air  temperature  of  25  °F, 
estimate  the  rate  of  change  of  the  windchill  tempera- 
ture with  respect  to  windspeed. 

15.  Windchill  tables  are  constructed  from  empirically  de- 
rived formulas  for  heat  loss  from  an  exposed  sur- 
face. Early  experimental  work  of  P.  A.  Siple  and  C.  F. 
Passel,4  resulted  in  the  following  formula: 

W  =  91 .4  +  (t  -  91 .4)(0.474  +  0.304V?  -  0.0203s). 


Here  W  denotes  windchill  temperature  (in  degrees 
Fahrenheit),  t  the  air  temperature  (for  t  <  9 1 .4  °F),  and 
s  the  windspeed  in  miles  per  hour  (for  s  >  4  mph).5 

(a)  Compare  your  answers  in  Exercises  1 1  and  12  with 
those  computed  directly  from  the  Siple  formula 
just  mentioned. 

(b)  Discuss  any  differences  you  observe  between  your 
answers  to  Exercises  1 1  and  12  and  your  answers 
to  part  (a). 

(c)  Why  is  it  necessary  to  take  t  <  91.4°F  and 
s  >  4  mph  in  the  Siple  formula?  (Don't  look  for 
a  purely  mathematical  reason;  think  about  the 
model.) 

1 6.  Recent  research  led  the  United  States  National  Weather 
Service  to  employ  a  new  formula  for  calculating  wind- 
chill values  beginning  November  1,  2001.  In  partic- 
ular, the  table  on  page  85  was  constructed  from  the 
formula 

W  =  35.74  +  0.621f  -  35.75j016  +  0.4275^016. 

Here,  as  in  the  Siple  formula  of  Exercise  15,  W  de- 
notes windchill  temperature  (in  degrees  Fahrenheit), 
t  the  air  temperature  (for  t  <  50 °F),  and  s  the  wind- 
speed  in  miles  per  hour  (for  s  >  3  mph).6  Compare 
your  answers  in  Exercises  13  and  14  with  those  com- 
puted directly  from  the  National  Weather  Service 
formula  above. 

17.  In  this  problem  you  will  compare  graphically  the  two 
windchill  formulas  given  in  Exercises  15  and  16. 

(a)  If  W\  (s,  t)  denotes  the  windchill  function  given  by 
the  Siple  formula  in  Exercise  15  and  0  the 
windchill  function  given  by  the  National  Weather 
Service  formula  in  Exercise  16,  graph  the  curves 
y  =  Wi(s,  40)  and  y  =  Wi(s,  40)  on  the  same  set 
of  axes.  (Let  s  vary  between  3  and  120  mph.)  In 
addition,  graph  other  pairs  of  curves  y  =  W\(s,  to), 
y  =  W2(s,  ?o)  for  other  values  of  to.  Discuss  what 
your  results  tell  you  about  the  two  windchill 
formulas. 

(b)  Now  graph  pairs  of  curves  y  =  Wi(sq,  t),  y  = 
W2CS0,  t)  for  various  constant  values  so  for  wind- 
speed.  Discuss  your  results. 

(c)  Finally,  graph  the  surfaces  z  =  W\{s,  t)  and 
z  =  W2(s,  t)  and  comment. 


"Measurements  of  dry  atmospheric  cooling  in  sub  freezing  temperatures,"  Proc.  Amer.  Phil.  Soc,  89 
(1945),  177-199. 

5  From  Bob  Rilling,  Atmospheric  Technology  Division,  National  Center  for  Atmospheric  Research 
(NCAR),  "Calculating  Windchill  Values,"  February  12,  1996.  Found  online  at  http://www.atd.ucar.edu/ 
homes/rilling/wc_formula.html  (July  31,  2010). 

6  NOAA,  National  Weather  Service,  Office  of  Climate,  Water,  and  Weather  Services,  "NWS  Wind  Chill 
Temperature  Index."  February  26,  2004.  <http://www.nws.noaa.gov/om/  windchill>  (July  31,  2010). 


1  86       Chapter  2  |  Differentiation  in  Several  Variables 


1 8.  Consider  the  sphere  of  radius  3  centered  at  the  origin. 
The  plane  tangent  to  the  sphere  at  (1,  2,  2)  intersects 
the  x-axis  at  a  point  P.  Find  the  coordinates  of  P. 

1 9.  Show  that  the  plane  tangent  to  a  sphere  at  a  point  P  on 
the  sphere  is  always  perpendicular  to  the  vector  OP 
from  the  center  O  of  the  sphere  to  P.  (Hint:  Locate  the 
sphere  so  its  center  is  at  the  origin  in  R3.) 

20.  The  surface  z  =  3x2  +  |x3  —  ^x4  —  4y2  is  inter- 
sected by  the  plane  2x  —  y  =  I.  The  resulting  intersec- 
tion is  a  curve  on  the  surface.  Find  a  set  of  parametric 
equations  for  the  line  tangent  to  this  curve  at  the  point 

21.  Consider  the  cone  z2  =  x2  +  y2. 

(a)  Find  an  equation  of  the  plane  tangent  to  the  cone 
at  the  point  (3,  -4,  5). 

(b)  Find  an  equation  of  the  plane  tangent  to  the  cone 
at  the  point  (a,  b,  c). 

(c)  Show  that  every  tangent  plane  to  the  cone  must 
pass  through  the  origin. 


22.  Show  that  the  two  surfaces 


Si:  z  =  xy     and     S2:  z  =  \x2  —  y 
intersect  perpendicularly  at  the  point  (2,  1,2). 

23.  Consider  the  surface  z  =  x2  +  Ay2. 

(a)  Find  an  equation  for  the  plane  that  is  tangent  to  the 
surface  at  the  point  (1,  —1,5). 

(b)  Now  suppose  that  the  surface  is  intersected  with  the 
plane  x  =  1 .  The  resulting  intersection  is  a  curve 
on  the  surface  (and  is  a  curve  in  the  plane  x  =  1 
as  well).  Give  a  set  of  parametric  equations  for  the 
line  in  R3  that  is  tangent  to  this  curve  at  the  point 
(1 ,  —  1 ,  5).  A  rough  sketch  may  help  your  thinking. 

24.  A  turtleneck  sweater  has  been  washed  and  is  now  tum- 
bling in  the  dryer,  along  with  the  rest  of  the  laundry.  At 
a  particular  moment  to,  the  neck  of  the  sweater  mea- 
sures 1 8  inches  in  circumference  and  3  inches  in  length. 
However,  the  sweater  is  100%  cotton,  so  that  at  to  the 
heat  of  the  dryer  is  causing  the  neck  circumference  to 
shrink  at  a  rate  of  0.2  in/min,  while  the  twisting  and 
tumbling  action  is  causing  the  length  of  the  neck  to 
stretch  at  the  rate  of  0.1  in/min.  How  is  the  volume  V 
of  the  space  inside  the  neck  changing  at  t  =  ?o?  Is  V 
increasing  or  decreasing  at  that  moment? 

25.  A  factory  generates  air  pollution  each  day  according 
to  the  formula 

P(S,  T)  =  33052/3T4/5, 

where  S  denotes  the  number  of  machine  stations  in 
operation  and  T  denotes  the  average  daily  tempera- 
ture. At  the  moment,  75  stations  are  in  regular  use  and 
the  average  daily  temperature  is  15  °C.  If  the  average 


temperature  is  rising  at  the  rate  of  0.2°C/day  and  the 
number  of  stations  being  used  is  falling  at  a  rate  of 
2  per  month,  at  what  rate  is  the  amount  of  pollution 
changing?  (Note:  Assume  that  there  are  24  workdays 
per  month.) 

26.  Economists  attempt  to  quantify  how  useful  or  satisfy- 
ing people  find  goods  or  services  by  means  of  utility 
functions.  Suppose  that  the  utility  a  particular  individ- 
ual derives  from  consuming  x  ounces  of  soda  per  week 
and  watching  y  minutes  of  television  per  week  is 


u(x,  y)  =  1 


-0.001.v2-0.00005v2 


Further  suppose  that  she  currently  drinks  80  oz  of  soda 
per  week  and  watches  240  min  of  TV  each  week.  If  she 
were  to  increase  her  soda  consumption  by  5  oz/week 
and  cut  back  on  her  TV  viewing  by  15  min/week,  is 
the  utility  she  derives  from  these  changes  increasing 
or  decreasing?  At  what  rate? 

27.  Suppose  that  w  =  x2  +  y2  +  z2  andx  =  p  cos#  s'mcp, 
y  =  p  sin  6  sin  cp,  z  =  p  cos  cp .  (Note  that  the  equations 
for  x,  y,  and  z  in  terms  of  p,  cp,  and  6  are  just  the  con- 
version relations  from  spherical  to  rectangular  coordi- 
nates.) 

(a)  Use  the  chain  rule  to  compute  dw/dp,  dw/d(p, 
and  dw/dd.  Simplify  your  answers  as  much  as 
possible. 

(b)  Substitute  p,  (p,  and  9  for  x,  y,  and  z  in  the  original 
expression  for  w.  Can  you  explain  your  answer  in 
part  (a)? 

28.  If  w  =  f  (—t-^)'  show  that 

,3w  ,3iu 

x2  y2 —  =  0. 

3x  dy 

(You  should  assume  that  /  is  a  differentiable  function 
of  one  variable.) 

29.  Let  z  =  g(x,  y)  be  a  function  of  class  C2,  and  let 
x  =  er  cos#,  v  =  er  sin#. 

(a)  Use  the  chain  rule  to  find  dz/dr  and  3 z/ 9 6  in  terms 
of  dz/dx  and  dz/dy.  Use  your  results  to  solve  for 
3z/3x  and  dz/dy  in  terms  of  dz/dr  and  dz/d0. 

(b)  Use  part  (a)  and  the  product  rule  to  show  that 


32z      32z  _  _2r  {dh  dh 

3x2  +  ^2  -  e       [dr2  +  d92 


30.  (a)  Use  the  function  /(x,  y)  =  xy  (=  eylnx)  and  the 

d 

multivariable  chain  rule  to  calculate  —  (w"). 

du 

(b)  Use  the  multivariable  chain  rule  to  calculate 

^((sinOcos'). 
dt 


Miscellaneous  Exercises  for  Chapter  2 


31.  Use  the  function  f(x,  y,  z)  =  xy  andthemultivariable 

d  ,  «, 

chain  rule  to  calculate  —  (u  ). 

du 

32.  Suppose  that  /:  R"  — »  R  is  a  function  of  class  C2.  The 
Laplacian  of  /,  denoted  V2/,  is  defined  to  be 


V2/ 


92/  a2/ 
— —  -i — —  +  ■ 

dx}  dx2 


+ 


a2/ 

dx2' 


When  n  =  2  or  3,  this  construction  is  important  when 
studying  certain  differential  equations  that  model  phys- 
ical phenomena,  such  as  the  heat  or  wave  equations. 
(See  Exercises  28  and  29  of  §2.4.)  Now  suppose  that 
/  depends  only  on  the  distance  x  =  (x\,  . . . ,  x„)  is 
from  the  origin  in  R" ;  that  is,  suppose  that  /(x)  =  g(r) 
for  some  function  g,  where  r  =  ||x|| .  Show  that  for  all 
x  /  0,  the  Laplacian  is  given  by 


V2/ 


g'(r)  +  g"(r). 


33.  (a)  Consider  a  function  f(x,  y)  of  class  C4.  Show 
that  if  we  apply  the  Laplacian  operator  V2  = 
d2/dx2  +  d2/dy2  twice  to  /,  we  obtain 


V2(V2/) 


 -L  +  2  -  

dx4  dx2dy2 


+ 


9V 
3y4  ' 


(b)  Now  suppose  that  /  is  a  function  of «  variables  of 
class  C4.  Show  that 


V2(V2/)=£ 


94/ 


,  dxfdx1- 

',7  =  1       '  J 

Functions  that  satisfy  the  partial  differential  equa- 
tion V2(V2/)  =  0  are  called  Inharmonic  func- 
tions and  arise  in  the  theoretical  study  of  elasticity. 

34.  Livinia,  the  housefly,  finds  herself  caught  in  the  oven 
at  the  point  (0,  0,  1).  The  temperature  at  points  in  the 
oven  is  given  by  the  function 

T(x,y,z)=  \0(xe->2  +ze~xl), 

where  the  units  are  in  degrees  Celsius. 

(a)  IfLivinia  begins  to  move  toward  the  point  (2,  3,  1), 
at  what  rate  (in  deg/cm)  does  she  find  the  temper- 
ature changing? 

(b)  In  what  direction  should  she  move  in  order  to  cool 
off  as  rapidly  as  possible? 

(c)  Suppose  that  Livinia  can  fly  at  a  speed  of  3  cm/sec. 
If  she  moves  in  the  direction  of  part  (b),  at  what 
(instantaneous)  rate  (in  deg/sec)  will  she  find  the 
temperature  to  be  changing? 

35.  Consider  the  surface  given  in  cylindrical  coordinates 
by  the  equation  z  =  r  cos  36. 

(a)  Describe  this  surface  in  Cartesian  coordinates,  that 
is,  asz  =  f{x,y). 


(b)  Is  /  continuous  at  the  origin?  (Hint:  Think  cylin- 
drical.) 

(c)  Find  expressions  for  df /dx  and  df/dy  at  points 
other  than  (0,  0).  Give  values  for  df /dx  and  df/  dy 
at  (0,  0)  by  looking  at  the  partial  functions  of  / 
through  (x,  0)  and  (0,  y)  and  taking  one-variable 
limits. 

(d)  Show  that  the  directional  derivative  Duf(0,  0)  ex- 
ists for  every  direction  (unit  vector)  u.  (Hint:  Think 
in  cylindrical  coordinates  again  and  note  that  you 
can  specify  a  direction  through  the  origin  in  the 
xy-plane  by  choosing  a  particular  constant  value 
for  9.) 

(e)  Show  directly  (by  examining  the  expression  for 
df/dy  when(jc,  y)  /  (0,  0)  and  also  using  part  (c)) 
that  df/dy  is  not  continuous  at  (0,  0). 

(f)  Sketch  the  graph  of  the  surface,  perhaps  using  a 
computer  to  do  so. 


36.  The  partial  differential  equation 

d2u      d2u  d2u 
Jx2  +  Jv2  +  Jz2  ~C~dt2 


d2u 


is  known  as  the  wave  equation.  It  models  the  motion 
of  a  wave  u  ( x ,  y ,  z ,  t )  in  R3  and  was  originally  derived 
by  Johann  Bernoulli  in  1727.  In  this  equation,  c  is  a 
positive  constant,  the  variables  x,  y,  and  z  represent 
spatial  coordinates,  and  the  variable  t  represents  time. 

(a)  Let  u  =  cos(x  —  t)  +  sin(x  +  t)  —  2ez+l  —  (y  — 
f)3  •  Show  that  u  satisfies  the  wave  equation  with 
c=  1. 

(b)  More  generally,  show  that  if  f\,  fa,  g\,  g%,  hi,  and 
hi  are  any  twice  differentiable  functions  of  a  single 
variable,  then 

u(x,  y,  z,  t)  =  fi(x  -  t)  +  f2(x  +  t) 

+  gi(y  - 1)  +  g2(y  + 1) 

+  hl(z-t)  +  h2(z  +  t) 

satisfies  the  wave  equation  with  c  =  1 . 

Let  X  be  an  open  set  in  R".  A  function  F:  X  — >  Ris  said  to  be 
homogeneous  of  degree  d  if,  for  all  x  =  (x\,  X2,  ■  ■  ■ ,  xn)  e  X 
and  all  t  €  R  such  that  tx  €  X,  we  have 

F(tx\ ,  tx2,  .  ■  ■ ,  tx„)  =  td F(x\ ,  x2,  . . . ,  xn). 

Exercises  37—44  concern  homogeneous  functions. 

In  Exercises  37-41,  which  of  the  given  functions  are  homoge- 
neous ?  For  those  that  are,  indicate  the  degree  dof  homogeneity. 

37.  F(x,y)  =  x3  +xy2  -  6y3 

38.  F(x,  y,  z)  =  x3y  -  x2z2  +  z& 

39.  F(x,  y,  z)  =  zy2  -  x3  +  x2z 


1  88       Chapter  2  |  Differentiation  in  Several  Variables 


40.  F(x,  y)  =  e 

41.  F(x,y,z) 


JC3  +  X2  v 


yz'- 


x\ 


z  +  Ixz2 


42.  If  F(x,  y,  z)  is  a  polynomial,  characterize  what  it 
means  to  say  that  F  is  homogeneous  of  degree  d  (i.e., 
explain  what  must  be  true  about  the  polynomial  if  it  is 
to  be  homogeneous  of  degree  d). 

43.  Suppose  F{x\,  X2,  . . . ,  xn)  is  differentiable  and  homo- 
geneous of  degree  d.  Prove  Euler's  formula: 

dF         dF  dF 
xi-  \-x2-  1  \-x„  —  =dF. 

OXl  0X2  OXn 


(Hint:  Take  the  equation  F(tx\,tX2,  ■  ■  ■  ,txn)  = 
td F(x\ ,  X2,  . ..  ,xn)  that  defines  homogeneity  and  dif- 
ferentiate with  respect  to  /  .) 

44.  Generalize  Euler's  formula  as  follows:  If  F  is  of  class 
C2  and  homogeneous  of  degree  d,  then 

d2F 
f-r1.  dxjdx: 

Can  you  conjecture  what  an  analogous  formula 
involving  the  fcth-order  partial  derivatives  should  look 
like? 


3.1  Parametrized  Curves  and 
Kepler's  Laws 

3.2  Arclength  and  Differential 
Geometry 

3.3  Vector  Fields:  An 
Introduction 

3.4  Gradient,  Divergence,  Curl, 
and  the  Del  Operator 

True/False  Exercises  for 
Chapter  3 

Miscellaneous  Exercises 
for  Chapter  3 


Z 


y 


X 


Figure  3.1  The  path  x  of 
Example  1. 


Vector-Valued 
Functions 


Introduction 

The  primary  focus  of  Chapter  2  was  on  scalar-valued  functions,  although  general 
mappings  from  R"  to  Rm  were  considered  occasionally.  This  chapter  concerns 
vector- valued  functions  of  two  special  types: 

1.  Continuous  mappings  of  one  variable  (i.e.,  functions  x:  /  C  R  -+  R",  where 
/  is  an  interval,  called  paths  in  R"). 

2.  Mappings  from  (subsets  of)  R"  to  itself  (called  vector  fields). 

An  understanding  of  both  concepts  is  required  later,  when  we  discuss  line  and 
surface  integrals. 

3.1   Parametrized  Curves  and  Kepler's  Laws 

Paths  in  Rn 

We  begin  with  a  simple  definition.  Let  /  denote  any  interval  in  R.  (So  /  can  be 
of  the  form  [a,  b],  (a,  b),  [a,  b),  {a,  b],  [a,  oo),  (a,  oo),  (— oo,  b],  (—00,  b),  or 
(—00,  00)  =  R.) 


DEFINITION  1.1    A  path  in  R"  is  a  continuous  function  x:  /  -»  R" .  If  /  = 

[a,  b]  for  some  numbers  a  <b,  then  the  points  x(a)  and  x(b)  are  called  the 
endpoints  of  the  path  x.  (Similar  definitions  apply  if  /  =  [a,  b),  [a,  00),  etc.) 


EXAMPLE  1  Let  a  and  b  be  vectors  in  R  with  a  /  0.  Then  the  function 
x:  (—00,  00)  ->  R3  given  by 

x(f)  =  b  +  ta 

defines  the  path  along  the  straight  line  parallel  to  a  and  passing  through  the  end- 
point  of  the  position  vector  of  b  as  in  Figure  3.1.  (See  formula  ( 1 )  of  §  1 .2.)  ♦ 

EXAMPLE  2   The  path  y:  [0,  2tt)  ->  R2  given  by 

y(f)  =  (3  cosf,  3  sinf) 

can  be  thought  of  as  the  path  of  a  particle  that  travels  once,  counterclockwise, 
around  a  circle  of  radius  3  (Figure  3.2).  ♦ 


190       Chapter  3  I  Vector-Valued  Functions 


Figure  3.4  The  path  x  and  its 
velocity  vector  v. 


EXAMPLE  3   The  map  z:  R  ->■  R3  defined  by 

z(f)  =  (a  cos  t,  a  sin?,  bt),     a,  b  constants  {a  >  0) 

is  called  a  circular  helix,  so  named  because  its  projection  in  the  xy-plane  is  a 
circle  of  radius  a.  The  helix  itself  lies  in  the  right  circular  cylinder  x2  +  y2  =  a2 
(Figure  3.3).  The  value  of  b  determines  how  tightly  the  helix  twists.  ♦ 

We  distinguish  between  a  path  x  and  its  range  or  image  set  x(7),  the  latter 
being  a  curve  in  R".  By  definition,  a  path  is  a  function,  a  dynamic  object  (at  least 
when  we  imagine  the  independent  variable  /  to  represent  time),  whereas  a  curve 
is  a  static  figure  in  space.  With  such  a  point  of  view,  it  is  natural  for  us  to  consider 
the  derivative  Dx(t ),  which  we  also  write  as  x'(/ )  or  v(/),  to  be  the  velocity  vector 
of  the  path.  We  can  readily  justify  such  terminology.  Since 

x(t)  =  (x1(t),x2(t),...,xn(t)) 


is  a  function  of  just  one  variable, 


v(/)  =  x'(t)  =  lim 


x(r  +  At)  -  x(t) 
At  ' 


Thus,  v(/)  is  the  instantaneous  rate  of  change  of  position  x(t)  with  respect  to  t 
(time),  so  it  can  appropriately  be  called  velocity.  Figure  3.4  provides  an  indication 
as  to  why  we  draw  v(f)  as  a  vector  tangent  to  the  path  at  x(f).  Continuing  in  this 
vein,  we  introduce  the  following  terminology: 


DEFINITION  1 .2  Let  x:  /  ->■  R"  be  a  differentiable  path.  Then  the  velocity 
\(t)  =  x'(t)  exists,  and  we  define  the  speed  of  x  to  be  the  magnitude  of 
velocity;  that  is, 

Speed  =  ||v(OI|. 

If  v  is  itself  differentiable,  then  we  call  V(t)  =  x"(t)  the  acceleration  of  x 
and  denote  it  by  m(t). 


EXAMPLE  4   The  helix  x(t )  =  (a  cos  t ,  a  sin  t ,  bt)  has 

v(/)  =  —  a  sinf  i  +  a  cost  j  +  fok    and    a(f)  =  —  a  cos/  i  —  a  sin/  j. 


3.1  ;  Parametrized  Curves  and  Kepler's  Laws 


Thus,  the  acceleration  vector  is  parallel  to  the  xy-plane  (i.e.,  is  horizontal).  The 
speed  of  this  helical  path  is 

||v(OII  =  -vA-a  sin t)2  +  (a  cos tf  +  b2  =  y/a2  +  b2, 
which  is  constant.  ♦ 

The  velocity  vector  v  is  important  for  another  reason,  namely,  for  finding  equations 
of  tangent  lines  to  paths.  The  tangent  line  to  a  differentiate  path  x,  at  the  point 
x0  =  x(?o),  is  the  line  through  xo  that  is  parallel  to  any  (nonzero)  tangent  vector 
to  x  at  xo.  Since  v(t),  when  nonzero,  is  always  tangent  to  x(t),  we  may  use  equa- 
tion ( 1 )  of  §  1 .2  to  obtain  the  following  vector  parametric  equation  for  the  tangent 
line: 

l(s)  =  Xo  +  SV().  (1) 

Here  vo  =  v(to)  and  s  may  be  any  real  number. 

In  equation  (1),  we  have  1(0)  =  Xo.  To  relate  the  new  parameter  s  to  the 
original  parameter  t  for  the  path,  we  set  s  =  t  —  to  and  establish  the  following 
result: 


PROPOSITION  1 .3  Let  x  be  a  differentiate  path  and  assume  that  v0  = 
v(?o)  ^  0.  Then  a  vector  parametric  equation  for  the  line  tangent  to  x  at  x0  =  x(t0) 
is  either 

l(s)  =  x0  +  s\0  (2) 

or 

1(0  =  x0  +  (t  -  f0)v0.  (3) 

(See  Figure  3.5.) 


EXAMPLE  5  If  x(?)  =  (it  +  2,  t2  -  7,  t  -  t2),  we  find  parametric  equations 
for  the  line  tangent  to  x  at  (5,  —  6,  0)  =  x(l). 

For  this  path,  v(0  =  x'(t)  =  3i  +  2/j  +  (1  -  2t)k,  so  that 


Thus,  by  formula  (3), 

1(0 


v0  =  v(l)  =  3i  +  2j-k. 


(5i-6j)  +  (f-  l)(3i  +  2j-k). 


Taking  components,  we  read  off  the  parametric  equations  for  the  coordinates 
of  the  tangent  line  as 


x  =  3t  +2 
y  =  2t  -  8 . 
z=\-t 


The  physical  significance  of  the  tangent  line  is  this:  Suppose  a  particle  of 
mass  m  travels  along  a  path  x.  If,  suddenly,  at  t  =  to,  all  forces  cease  to  act  on  the 
particle  (so  that,  by  Newton's  second  law  of  motion  F  =  ma,  we  have  a(/)  =  0 
for  t  >  to),  then  the  particle  will  follow  the  tangent  line  path  of  equation  (3). 


192       Chapter  3  I  Vector-Valued  Functions 


EXAMPLE  6  If  Roger  Ramjet  is  fired  from  a  camion,  then  we  can  use  vectors 
to  describe  his  trajectory.  (See  Figure  3.6.) 

y 


Roger's  path 


x 


Figure  3.6  Roger  Ramjet's  path. 

We'll  assume  that  Roger  is  given  an  initial  velocity  vector  vo  by  virtue  of  the 
firing  of  the  cannon  and  that  thereafter  the  only  force  acting  on  Roger  is  due  to 
gravity  (so,  in  particular,  we  neglect  any  air  resistance).  Let  us  choose  coordinates 
so  that  Roger  is  initially  at  the  origin,  and  throughout  our  calculations  we  '11  neglect 
the  height  of  the  cannon.  Let  x(r)  =  (x(f),  y(t))  denote  Roger's  path.  Then  the 
information  we  have  is 

a(t)  =  x"(t)=-g\ 
(i.e.,  the  acceleration  due  to  gravity  is  constant  and  points  downward);  hence, 

v(0)  =  x'(0)  =  v0 

and 


x(0)  =  0. 

Since  a(/)  =  V(t),  we  simply  integrate  the  expression  for  acceleration  compo- 
nentwise to  find  the  velocity: 


v(0 


=  j  a(t)dt  =  j  -g]dr  =  -gtj  + 


Here  c  is  an  arbitrary  constant  vector  (the  "constant  of  integration").  Since  v(0)  = 
vo,  we  must  have  c  =  vo,  so  that 

v(0  =  -gtj  +  v0. 
Integrating  again  to  find  the  path, 

x(t)  =  j  v(t)dt  =  j  (-gt\  +  \0)dt  =  -^gt2]  +  t\0  +  d, 

where  d  is  another  arbitrary  constant  vector.  From  the  remaining  fact  that  x(0)  =  0, 
we  conclude  that 

1  7 

x(0  =  -^gn  +  ty0  (4) 

describes  Roger's  path. 

To  understand  equation  (4)  better,  we  write  vo  in  terms  of  its  components: 

vo  =  vo  cos  9  i  +  vo  sin  9  j. 

Here  vo  =  ||vo||  is  the  initial  speed.  (We're  really  doing  nothing  more  than 
expressing  the  rectangular  components  of  vo  in  terms  of  polar  coordinates. 


3.1  :  Parametrized  Curves  and  Kepler's  Laws 


See  Figure  3.7.)  Thus, 


x(r)  =  —  \gt2\  +  f(i>0cos#i  +  vosin#  j) 


=  (v0  cos  9)t\  +  I  (u0  sin  9)t  —  -gt2 


1 

2 

From  this,  we  may  read  off  the  parametric  equations: 

x  =  (vo  cos6)t 

1    ,  - 
y  =  (v0  sinfl)/  -  -gtz 

from  which  it  is  not  difficult  to  check  that  Roger's  path  traces  a  parabola.  ♦ 

Here  are  two  practical  questions  concerning  the  set-up  of  Example  6:  First,  for 
a  given  initial  velocity,  how  far  does  Roger  travel  horizontally?  Second  for  a  given 
initial  speed  how  should  the  cannon  be  aimed  so  that  Roger  travels  (horizontally) 
as  far  as  possible?  To  find  the  range  of  the  cannon  shot  and  thereby  answer  the 
first  question,  we  need  to  know  when  y  =  0  (i.e.,  when  Roger  hits  the  ground). 
Thus,  we  solve 

(u0  sin 6>)r  -  \gt2  =  t(v0sm9  -  \gt)  =  0 

for  t.  Hence,  y  =  0  when  t  =  0  (which  is  when  Roger  blasts  off)  and  when 
t  =  (2vq  sin9)/g.  At  this  later  time, 

(2vo  sin#\          sin  29 
—   =  ■  (5) 

8      J  S 

Formula  (5)  is  Roger's  horizontal  range  for  a  given  initial  velocity.  To  maximize 
the  range  for  a  given  initial  speed  vo,  we  must  choose  9  so  that  (v2  sin29)/g  is 
as  large  as  possible.  Clearly,  this  happens  when  sin  29  =  1  (i.e.,  when  9  =  jt/4). 


Kepler's  Laws  of  Planetary  Motion  (optional)   

Since  classical  antiquity,  individuals  have  sought  to  understand  the  motions  of 
the  planets  and  stars.  The  majority  of  the  ancient  astronomers,  using  a  combina- 
tion of  crude  observation  and  faith,  believed  all  heavenly  bodies  revolved  around 
the  earth.  Fortunately,  the  heliocentric  (or  "sun-centered")  theory  of  Nicholas 
Copernicus  (1473-1543)  did  eventually  gain  favor  as  observational  techniques 
improved.  However,  it  was  still  believed  that  the  planets  traveled  in  circular  or- 
bits around  the  sun.  This  circular  orbit  theory  did  not  correctly  predict  planetary 
positions,  so  astronomers  postulated  the  existence  of  epicycles,  smaller  circular 
orbits  traveling  along  the  major  circular  arc,  an  example  of  which  is  shown  in 
Figure  3.8.  Although  positional  calculations  with  epicycles  yielded  results  closer 
to  the  observed  data,  they  still  were  not  correct.  Attempts  at  further  improvements 
were  made  using  second-  and  third-order  epicycles,  but  any  gains  in  predictive 
power  were  made  at  a  cost  of  considerable  calculational  complexity.  A  new  idea 
was  needed.  Such  inspiration  came  from  Johannes  Kepler  (1571-1630),  son  of  a 
saloonkeeper  and  assistant  to  the  Danish  astronomer  Tycho  Brahe.  The  classical 
astronomers  were  "stuck  on  circles"  for  they  believed  the  circle  to  be  a  perfect 
form  and  that  God  would  use  only  such  perfect  figures  for  planetary  motion. 
Kepler,  however,  considered  the  other  conic  sections  to  be  as  elegant  as  the  cir- 
cle and  so  hypothesized  the  simple  theory  that  planetary  orbits  are  elliptical. 
Empirical  evidence  bore  out  this  theory. 


194       Chapter  3  |  Vector-Valued  Functions 


Figure  3.9  Kepler's  second  law 
of  planetary  motion:  If 
t2  —  t\  =  t4  —  f3,  then  A\  =  A2, 
where  A;  and  A2  are  the  areas  of 
the  shaded  regions. 


Kepler's  three  laws  of  planetary  motion  are 

1.  The  orbit  of  a  planet  is  elliptical,  with  the  sun  at  a  focus  of  the  ellipse. 

2.  During  equal  periods  of  time,  a  planet  sweeps  through  equal  areas  with  respect 
to  the  sun.  (See  Figure  3.9.) 

3.  The  square  of  the  period  of  one  elliptical  orbit  is  proportional  to  the  cube  of 
the  length  of  the  semimajor  axis  of  the  ellipse. 

Kepler's  laws  changed  the  face  of  astronomy.  We  emphasize,  however,  that 
they  were  discovered  empirically,  not  analytically  derived  from  general  physical 
laws.  The  first  analytic  derivation  is  frequently  credited  to  Newton,  who  claimed 
to  have  established  Kepler's  laws  (at  least  the  first  and  third  laws)  in  Book  I  of 
his  Philosophiae  Naturalis  Principia  Mathematica  (1687).  However,  a  number  of 
scientists  and  historians  of  science  now  consider  Newton's  proof  of  Kepler's  first 
law  to  be  flawed  and  that  Johann  Bernoulli  (1667-1748)  offered  the  first  rigorous 
derivation  in  1710.1  In  the  discussion  that  follows,  Newton's  law  of  universal 
gravitation  is  used  to  prove  all  three  of  Kepler's  laws. 

In  our  work  below,  we  assume  that  the  only  physical  effects  are  those  be- 
tween the  sun  and  a  single  planet — the  so-called  two-body  problem.  (The  n-body 
problem,  where  n  >  3  is,  by  contrast,  an  important  area  of  current  mathematical 
research.)  To  set  the  stage  for  our  calculations,  we  take  the  sun  to  be  fixed  at  the 
origin  O  in  R3  and  the  planet  to  be  at  the  moving  position  P.  We  also  need  the 
following  two  "vector  product  rules,"  whose  proofs  we  leave  to  you: 


PROPOSITION  1.4 

1.  If  x  and  y  are  differentiable  paths  in  R",  then 

d                dx  dy 
—  (x  •  y)  =  v  •  h  x  •  — . 

dty  J   dt  dt 

2.  If  x  and  y  are  differentiable  paths  in  R3,  then 

d  dx  dy 

—  (x  xy)=  —  xy  +  xx  — . 
dty      J      dt     J  dt 


First,  we  establish  the  following  preliminary  result: 


PROPOSITION  1 .5  The  motion  of  the  planet  is  planar,  and  the  sun  lies  in  the 
planet's  plane  of  motion. 


PROOF  Let  r  =  OP.  Then  r  is  a  vector  whose  representative  arrow  has  its  tail 
fixed  at  O.  (Note  that  r  =  r(t);  that  is,  r  is  a  function  of  time.)  If  v  =  r'(f),  we 
will  show  that  r  x  v  is  a  constant  vector  c.  This  result,  in  turn,  implies  that  r  must 
always  be  perpendicular  to  c  and,  hence,  that  r  always  lies  in  a  plane  with  c  as 
normal  vector. 

To  show  that  r  x  v  is  constant,  we  show  that  its  derivative  is  zero.  By  part  2 
of  Proposition  1.4, 

d  dr  dy 

—  (r  xv)=  —  xv  +  rx  —  =  vxv+rxa, 

dt  dt  dt 


For  an  indication  of  the  more  recent  controversy  surrounding  Newton's  mathematical  accomplishments, 
see  R.  Weinstock,  "Isaac  Newton:  Credit  where  credit  won't  do,"  The  College  Mathematics  Journal,  25 
(1994),  no.  3,  179-192,  andC.  Wilson,  "Newton's  orbit  problem:  A  historian's  response,"  Ibid.,  193-200, 
and  related  papers. 


3.1  i  Parametrized  Curves  and  Kepler's  Laws 


by  the  definitions  of  velocity  and  acceleration.  We  know  that  vxv  =  0  (why?),  so 

d 

— (r  x  v)  =  r  x  a.  (6) 
dt 

Now  we  use  Newton's  laws.  Newton's  law  of  gravitation  tells  us  that  the  planet 
is  attracted  to  the  sun  with  a  force 

GMm 

F  =  —  u,  (7) 

where  G  is  Newton's  gravitational  constant  (=  6.6720  x  10~n  Nm2/kg2),  M 
is  the  mass  of  the  sun,  m  is  the  mass  of  the  planet  (in  kilograms),  r  =  ||r||,  and 
u  =  r/ 1|  i* ||  (distances  in  meters).  On  the  other  hand  Newton's  second  law  of 
motion  states  that,  for  the  planet, 

F  =  ma. 

Thus, 

GMm 
ma  =  r —  u, 

rL 

or 

GM 

a  =  -r.  (8) 

Therefore,  a  is  just  a  scalar  multiple  of  r  and  hence  is  always  parallel  to  r.  In 
view  of  equations  (6)  and  (8),  we  conclude  that 


d 

— (r  xv)  =  rxa  =  0 

dt 


(i.e.,  that  r  x  v  is  constant). 


THEOREM  1 .6  (Kepler's  first  law)  In  a  two-body  system  consisting  of  one 
sun  and  one  planet,  the  planet's  orbit  is  an  ellipse  and  the  sun  lies  at  one  focus  of 
that  ellipse. 


PROOF  We  will  eventually  find  a  polar  equation  for  the  planet's  orbit  and  see 
that  this  equation  defines  an  ellipse  as  described.  We  retain  the  notation  from 
the  proof  of  Proposition  1.5  and  take  coordinates  for  R3  so  that  the  sun  is  at  the 
origin,  and  the  path  of  the  planet  lies  in  the  xy-plane.  Then  the  constant  vector 
c  =  r  x  v  used  in  the  proof  of  Proposition  1.5  may  be  written  as  ck,  where  c  is 
some  nonzero  real  number.  This  set-up  is  shown  in  Figure  3.10. 

z 

c  =  r  x  v 


Figure  3.10  Establishing  Kepler's  laws. 


196       Chapter  3  I  Vector-Valued  Functions 


Step  1.  We  find  another  expression  for  c.  By  definition  of  u  in  formula  (7), 
r  =  ru,  so  that,  by  the  product  rule, 


d  du  dr 

v  =  —  (ru)  =  r—  +  —  u. 

dt  dt  dt 


Hence, 


/  du     dr    \       2  /      du\  dr 
c  =  r  x  v  =  (ru)  x    r  1  u    =r     ux  —    +  r — (u  x  u). 

\  dt      dt    J         \       dt  J  dt 

Since  uxu  must  be  zero,  we  conclude  that 

c  =  r'(uxf).  ,9, 

Step  2.  We  derive  the  polar  equation  for  the  orbit.  Before  doing  so,  however, 
note  the  following  result,  whose  proof  is  left  to  you  as  an  exercise: 


PROPOSITION  1.7  If  x(r)  has  constant  length  (i.e.,  ||x(f)||  is  constant  for 
all  r),  then  x  is  perpendicular  to  its  derivative  dx/dt. 


Continuing  now  with  the  main  argument,  note  that  the  vector  r(r)  is  denned 
so  that  its  magnitude  is  precisely  the  polar  coordinate  r  of  the  planet's  position. 
Using  equations  (8)  and  (9),  we  find  that 


axe 


GM 


u    x  r     u  x 


du 

dt 


=  -GM 


=  GM 


duX 

U  X  |  U  X   

dt  ) 


du\ 

U  X    X  u 

dt  J 


=  GM 


=  GM 


(u-u) 


du 

dt 


du 
u-  —  |  u 

dt 


du 
1  Ou 

dt 


(see  Exercise  27  of  §  1 .4) 
(by  Proposition  1.7) 


=  -  (GMu), 
dt 

since  G  and  M  are  constant.  On  the  other  hand,  we  can  "reverse"  the  product  rule 
to  find  that 

d\ 

a  x  c  =  —  x  c 
dt 

dy  dc 
=  —  xc  +  vx—      (since  c  is  constant) 
dt  dt 


=  —  (v  X  c). 

dt 


3.1  ;  Parametrized  Curves  and  Kepler's  Laws 


Figure  3.1 1  The  angle  9  is 
the  angle  between  r  and  d. 


Thus, 


and,  hence, 


d  d 
a  x  c  =  — (GMu)  =  — (v  x  c), 
dt  dt 


v  x  c  =  GMu  +  d, 


(10) 

where  d  is  an  arbitrary  constant  vector.  Because  both  v  x  c  and  u  lie  in  the  xy- 
plane,  so  must  d. 

Let  us  adjust  coordinates,  if  necessary,  so  that  d  points  in  the  i-direction  (i.e., 
so  that  d  =  di  for  some  d  e  R).  This  can  be  accomplished  by  rotating  the  whole 
set-up  about  the  z-axis,  which  does  not  lift  anything  lying  in  the  xy-plane  out  of 
that  plane.  Then  the  angle  between  r  (and  hence  u)  and  d  is  the  polar  angle  6  as 
shown  in  Figure  3.11. 

By  Theorem  3.3  of  Chapter  1, 

u-d=  ||u| 

Since  c  =  ||c||, 

2 


Idll  cos#  =  dcosi 


(11) 


c  •  c 

=  (r  x  v)  •  c 
=  r  •  (v  x  c) 
=  ru-(GMu  +  d) 


Hence, 


(Why?  See  formula  (4)  of  §1.4.) 
by  equation  (10). 

GMr  +  rd  cos6» 


by  equation  (11).  We  can  readily  solve  this  equation  for  r  to  obtain 

c2 

r  = 


(12) 


GM  +  d  cosO 
the  polar  equation  for  the  planet's  orbit. 

Step  3.  We  now  check  that  equation  (12)  really  does  define  an  ellipse  by 
converting  to  Cartesian  coordinates.  First,  we'll  rewrite  the  equation  as 

c2  (c2/GM) 
GM  +  d  cosO  ~  l  +  (d/GM)cose' 

and  then  let  p  =  c2/ GM,  e  =  d/ GM  for  convenience.  (Note  that  p  >  0.)  Hence, 
equation  (12)  becomes 

r  =  ■  (13) 

1  +  e  cos  e 

A  little  algebra  provides  the  equivalent  equation, 

r  =  p  —  er  cos6.  (14) 

Now  r  cose  =  x  (x  being  the  usual  Cartesian  coordinate),  so  that  equation  (14) 
is  equivalent  to 

r  =  p  —  ex. 

To  complete  the  conversion,  we  square  both  sides  and  find,  by  virtue  of  the  fact 
that  r2  =  x2  +  y2, 


x2  +  y2  =  p2 


2pex  +  e2x2 . 


198       Chapter  3  I  Vector-Valued  Functions 


A  little  more  algebra  reveals  that 

(1-  e2)x2  +  2pex  +  y2  =  p2.  (15) 

Therefore,  the  curve  described  by  the  preceding  equation  is  an  ellipse  if  0  < 
\e\  <  1,  a  parabola  if  e  =  ±1,  and  a  hyperbola  if  \e \  >  1 .  Analytically,  there  is  no 
way  to  eliminate  the  last  two  possibilities.  Indeed,  "uncaptured"  objects  such  as 
comets  or  expendable  deep  space  probes  can  have  hyperbolic  or  parabolic  orbits. 
However,  to  have  a  closed  orbit  (so  that  the  planet  repeats  its  transit  across  the 
sky),  we  are  forced  to  conclude  that  the  orbit  must  be  elliptical. 

More  can  be  said  about  the  elliptical  orbit.  Dividing  equation  (15)  by  1  —  e2 
and  completing  the  square  in  x,  we  have 

\  2  2  2 

x  + 


1  -e2      (1  -e2)2' 
This  is  equivalent  to  the  rather  awkward-looking  equation 

(x  +  pe/(l-e2)f  y2 

p2/(l  -  e2)2  p2/{\  -  e2)  y  J 

From  equation  (16),  we  see  that  the  ellipse  is  centered  at  the  point  (— pe/(l  —  e2), 
0),  that  its  semimajor  axis  has  length  a  =  p/(l  —  e2),  and  that  its  semiminor  axis 
has  length  b  =  p/^/l  —  e2.  The  foci  of  the  ellipse  are  at  a  distance 


yV  -  b2 


P2  P2  P\e\ 


(1  -  e2)2      1  -  e2 


from  the  center.  (See  Figure  3.12.)  Hence,  we  see  that  one  focus  must  be  at  the 
origin,  the  location  of  the  sun.  Our  proof  is,  therefore,  complete.  ■ 

Fortunately,  all  the  toil  involved  in  proving  the  first  law  will  pay  off  in  proofs 
of  the  second  and  third  laws,  which  are  considerably  shorter.  Again,  we  retain  all 
the  notation  we  already  introduced. 


THEOREM  1 .8  (Kepler's  second  law)  During  equal  intervals  of  time,  aplanet 
sweeps  through  equal  areas  with  respect  to  the  sun. 


Figure  3.1 2  The  ellipse  of  equation  (16). 


3.1  ;  Parametrized  Curves  and  Kepler's  Laws 


Po(r0,  B0) 


P(r,  6) 


Figure  3.1 3  The  shaded  area  A(9)  is 
given  by \r2  dip. 


PROOF  Fix  one  point  Pq  on  the  planet's  orbit.  Then  the  area  A  swept  between 
Po  and  a  second  (moving)  point  P  on  the  orbit  is  given  by  the  polar  area  integral 


A{6) 


f6  1  2 

L  2r 


dip. 


(See  Figure  3.13.)  Thus,  we  may  reformulate  Kepler's  law  to  say  that  dA/dt  is 
constant.  We  establish  this  reformulation  by  relating  dA/dt  to  a  known  constant, 
namely,  the  vector  c  =  r  x  v. 

By  the  chain  rule  (in  one  variable), 

dA  _  dA  d9 

dt       d6  dt ' 

By  the  fundamental  theorem  of  calculus, 

dA       d    f9  1  ,  ,        1  r  ,~ 


Hence, 


dA  _  1  2  dO 
dt  ~  2r  dt' 


(17) 


Now,  we  relate  c  to  dO/dt  by  means  of  equation  (9).  Therefore,  we  compute 

u  x  du/dt  in  terms  of  0.  Recall  that  u  =  -  r  and  r  =  r  cos  0  i  +  r  sin 6*  j.  Thus, 

r 

cos  0  i  +  sin  0  j 


u 

du 

dt  dt    '  dt 

Hence,  it  follows  by  direct  calculation  of  the  cross  product  that 


dO  dO 
smtf  —  i  +  cosO  —  j. 


c  =  r    u  x 


du 

dt 


2d0 
=  r  —  k, 
dt 


so  c  =  ||c||  =  r2d0/dt,  and  equation  (17)  implies  that 

dA  _  1 

~dt  ~  2C' 

a  constant. 


(18) 


THEOREM  1 .9  (Kepler's  third  law)  If  T  is  the  length  of  time  for  one  plane- 
tary orbit,  and  a  is  the  length  of  the  semimajor  axis  of  this  orbit,  then  T2  =  Ka3 
for  some  constant  K. 


Chapter  3  |  Vector-Valued  Functions 


3.1  Exercises 


PROOF  We  focus  on  the  total  area  enclosed  by  the  elliptical  orbit.  The  area  of  an 
ellipse  whose  semimajor  and  semiminor  axes  have  lengths  a  and  b,  respectively, 
is  nab.  This  area  must  also  be  that  swept  by  the  planet  in  the  time  interval  [0,  T]. 
Thus,  we  have 

'T  dA 
—  dt 
dt 

T  1 

-cdt  by  equation  ( 1 8) 


nab  =  I 

Jo 

-L 


Hence, 


1 

=  2CT- 

T  =  2-^,     so    r2=4"W.  (19) 

c  c2 

Now,  b  and  c  are  related  to  a,  so  these  quantities  must  be  replaced  before  we  are 
done.  In  particular,  from  equation  (16),  b2  =  p2/(l  —  e2),  so 


b2  =  pa. 


Also 

P 


c2 


GM 

(See  equations  (12)  and  (13).)  With  these  substitutions,  the  result  in  (19)  becomes 
t2  =  An2a2{pa)  =  /4^r2\  ^ 


pGM  \Gm) 


This  last  equation  shows  that  T2  is  proportional  to  a3,  but  it  says  even  more: 
The  constant  of  proportionality  4n2/GM  depends  entirely  on  the  mass  of  the 
sun — the  constant  is  the  same  for  any  planet  that  might  revolve  around  the  sun. 


In  Exercises  1-6,  sketch  the  images  of  the  following  paths,  us- 
ing arrows  to  indicate  the  direction  in  which  the  parameter 
increases: 

.    \x  =  It  -  1 

1.  {       ,     ,  ,     -1  <  t  <  1 
[y  =  3  -  t 

2.  x(t)  =  e'  i  +  <?"'  j 

_    \x  =  t  cos  t         ,  , 

3.  {  .     ,     -6n  <  t  <  6jt 
I  y  =  t  sin  t  ~  ~ 

.    \x  =  3  cost       „  - 
I  y  =  2  sin  2t         ~  ~ 

5.  x(r)  =  (f,3r2+l,0) 

6.  x(t)  =  (t,  t2,  t3) 

Calculate  the  velocity,  speed,  and  acceleration  of  the  paths 
given  in  Exercises  7—10. 

7.  x(r)  =  (3f-5)i  +  (2?  +  7)j 


8.  x(f )  =  5  cos  t  i  +  3  sin  t  j 

9.  x(r)  =  (t  sin*,  t  cos  t,  t2) 

10.  x(?)  =  (e',e2',2e') 

In  Exercises  11—14,  (a)  use  a  computer  to  give  a  plot  of  the 
given  path  x  over  the  indicated  interval  for  t;  identify  the  di- 
rection in  which  t  increases,  (b)  Show  that  the  path  lies  on  the 
given  surface  S. 

^  11.  x(?)  =  (3cosjrf,4sin7r?,2?),  -4  <  t  <  4;  S  is  ellip- 

x2  y2 

tical  cylinder  \-  - —  =  1 . 

9  16 

^12.  x(f)  =  (t  cost,  t  shU,  t),  -20  <  t  <  20;  S  is  cone 

Z2  =  x2  +  y2. 

^  13.  x(?)  =  (t  sin2r,  t  cos2r,  t2),  -6  <  t  <  6;  S  is  para- 
boloid z  =  x2  +  y2. 


3.1  |  Exercises  201 


^  14.  x(r)  =  (2cosf,  2sinr,  3  sin8f),  0  <  t  <  2jt;  S  is  cy- 
linders2 +  y2  =  4. 

In  Exercises  1 5-18,  find  an  equation  for  the  line  tangent  to  the 
given  path  at  the  indicated  value  for  the  parameter. 

15.  x(r)  =  te~>  i  +  e3'  j,  t  =  0 

16.  x(t)  =  4cos?i—  3sinr j  +  5tk,  t  =  tt/3 

17.  x(t)  =  (t2,t\t5),  t  =  2 

18.  x(r)  =  (cos(e'),  3  -  t2,  t),  t  =  1 

19.  (a)  Sketch  the  path  x(f)  =  (t,  f3  -  It  +  1). 

(b)  Calculate  the  line  tangent  to  x  when  t  =  2. 

(c)  Describe  the  image  of  x  by  an  equation  of  the  form 
y  =  f(x)  by  eliminating  t . 

(d)  Verify  your  answer  in  part  (b)  by  recalculating  the 
tangent  line,  using  your  result  in  part  (c). 

Exercises  20-23  concern  Roger  Ramjet  and  his  trajectory  when 
he  is  shot  from  a  cannon  as  in  Example  6  of  this  section. 

20.  Verify  that  Roger  Ramjet's  path  in  Example  6  is  indeed 
a  parabola. 

21 .  Suppose  that  Roger  is  fired  from  the  cannon  with  an 
angle  of  inclination  6  of  60°  and  an  initial  speed  Vq  of 
100  ft/sec.  What  is  the  maximum  height  Roger  attains? 

22.  Suppose  that  Roger  is  fired  from  the  cannon  with  an  an- 
gle of  inclination  6  of  60°  and  that  he  hits  the  ground 
1/2  mile  from  the  cannon.  What,  then,  was  Roger's 
initial  speed? 

23.  If  Roger  is  fired  from  the  cannon  with  an  initial  speed  of 
250  ft/sec,  what  angle  of  inclination  6  should  be  used 
so  that  Roger  hits  the  ground  1 500  ft  from  the  cannon? 

24.  Gertrude  is  aiming  a  Super  Drencher  water  pistol  at 
Egbert,  who  is  1.6  m  tall  and  is  standing  5  m  away. 
Gertrude  holds  the  water  gun  1  m  above  ground  at  an 
angle  a  of  elevation.  (See  Figure  3.14.) 

(a)  If  the  water  pistol  fires  with  an  initial  speed  of 
7  m/sec  and  an  elevation  angle  of  45°,  does  Egbert 
get  wet? 


(b)  If  the  water  pistol  fires  with  an  initial  speed  of 
8  m/sec,  what  possible  angles  of  elevation  will 
cause  Egbert  to  get  wet?  (Note:  You  will  want  to 
use  a  computer  algebra  system  or  a  graphics  cal- 
culator for  this  part.) 

25.  A  malfunctioning  rocket  is  traveling  according  to  the 
pathx(r)  =  (e2',  3f3  —  2t,  t  -  j)  in  the  hope  of  reach- 
ing a  repair  station  at  the  point  (7e4,  35,  5).  (Here 
t  represents  time  in  minutes  and  spatial  coordinates 
are  measured  in  miles.)  Alt  =  2,  the  rocket's  engines 
suddenly  cease.  Will  the  rocket  coast  into  the  repair 
station? 

26.  Two  billiard  balls  are  moving  on  a  (coordina- 
tized)  pool  table  according  to  the  respective  paths 


x(f)  =  (t 


!,  j  -  1J  and  y(f)  =  (t,  5  -  r),  where 
t  represents  time  measured  in  seconds. 

(a)  When  and  where  do  the  balls  collide? 

(b)  What  is  the  angle  formed  by  the  paths  of  the  balls 
at  the  collision  point? 

27.  Establish  part  1  of  Proposition  1.4  in  this  section:  If  x 
and  y  are  differentiable  paths  in  R" ,  show  that 


dt 


(x-y): 


dx 


dy 
dt ' 


28.  Establish  part  2  of  Proposition  1.4  in  this  section:  If  x 
and  y  are  differentiable  paths  in  R3,  show  that 


dt 


dx  dy 
(xxy)  =  —  xy  +  xx  — . 

dt  at 


29.  Prove  Proposition  1.7. 

30.  (a)  Show  that  the  path  x(f)  =  (cos  t,  cos  t  sin  t,  sin2  t) 

lies  on  a  unit  sphere. 

(b)  Verify  that  x(r)  is  always  perpendicular  to  the  ve- 
locity vector  v(f ). 

(c)  Use  Proposition  1 .7  to  show  that  if  a  differentiable 
path  lies  on  a  sphere  centered  at  the  origin,  then 
its  position  vector  is  always  perpendicular  to  its 
velocity  vector. 


Figure  3.14  Figure  for  Exercise  24. 


Chapter  3  |  Vector-Valued  Functions 


31.  Consider  the  path 

x  =  (a  +  b  cos  cot )  cos  t 
■  y  =  (a  +  b  cos  cot)  sin  t , 
z  =  b  sin  cot 

where  a,  b,  and  co  are  positive  constants  and  a  >  b. 

(a)  Use  a  computer  to  plot  this  path  when 

i.  a  =  3,  b  =  1,  and  co  =  15. 

ii.  a  =  5,  b  =  1,  and  co  =  15. 

iii.  a  =  5,  b  =  1,  and  w  =  25. 

Comment  on  how  the  values  of  a,  b,  and  <w  affect 
the  shapes  of  the  image  curves. 

(b)  Show  that  the  image  curve  lies  on  the  torus 

(vV  +  y2-a)2  +  z2  =  b2. 

(A  torus  is  the  surface  of  a  doughnut.) 

32.  For  the  path  x(f )  =  (e'  cos  t,  e'  sin  f ),  show  that  the  an- 
gle between  x(t)  and  x'(t)  remains  constant.  What  is 
the  angle? 

33.  Consider  the  path  x:  R      R2,  x(f )  =  (t2,  t3  -  t). 

(a)  Show  that  this  path  intersects  itself,  that  is,  that 
there  are  numbers  t\  and  ti  such  that  x(f i )  =  x(?2). 

(b)  At  the  point  where  the  path  intersects  itself,  it 
makes  sense  to  say  that  the  image  curve  has  two 
tangent  lines.  What  is  the  angle  between  these  tan- 
gent lines? 

34.  Although  the  path  x  :  [0,  2tt]  — >  R2,  x(f )  = 
(cosf,  sinf)  may  be  the  most  familiar  way  to  give  a 
parametric  description  of  a  unit  circle,  in  this  problem 
you  will  develop  a  different  set  of  parametric  equations 
that  gives  the  x-  and  v-coordinates  of  a  point  on  the 
circle  in  terms  of  rational  functions  of  the  parameter. 
(This  particular  parametrization  turns  out  to  be  useful 
in  the  branch  of  mathematics  known  as  number  theory.) 

To  set  things  up,  begin  with  the  unit  circle  x1  + 
y2  =  1  and  consider  all  lines  through  the  point  (—1,0). 
(See  Figure  3.15.)  Note  that  every  line  other  than  the 


vertical  line  x  =  —  1  intersects  the  circle  at  a  point 
(x,  y)  other  than  (—1,  0).  Let  the  parameter  t  be  the 
slope  of  the  line  joining  (—1,0)  and  a  point  (x,  y)  on 
the  circle. 


y 


(-1,0)1 

2"  Slope? 

 X 

Figure  3.15  Figure  for  Exercise  34. 


(a)  Give  an  equation  for  the  line  of  slope  t  joining 
(—1,0)  and  (x,  y).  (Your  answer  should  involve 
x,  y,  and  t.) 

(b)  Use  your  answer  in  part  (a)  to  write  y  in  terms  of 
x  and  t.  Then  substitute  this  expression  for  y  into 
the  equation  for  the  unit  circle.  Solve  the  resulting 
equations  for  x  in  terms  of  /.  Your  answer(s)  for  x 
will  give  the  points  of  intersection  of  the  line  and 
the  circle. 

(c)  Use  your  result  in  part  (b)  to  give  a  set  of  paramet- 
ric equations  for  points  (x,  y)  on  the  unit  circle. 

(d)  Does  your  parametrization  in  part  (c)  cover  the 
entire  circle?  Which,  if  any,  points  are  missed? 

35.  Let  x(f)  be  a  path  of  class  C1  that  does  not  pass  through 
the  origin  in  R3.  If  x(fo)  is  the  point  on  the  image  of  x 
closest  to  the  origin  and  x'(fo)  /  0,  show  that  the  po- 
sition vector  x(?o)  is  orthogonal  to  the  velocity  vector 
x'(fo). 


3.2  Arclength  and  Differential  Geometry 

In  this  section,  we  continue  our  general  study  of  parametrized  curves  in  R3, 
considering  how  to  measure  such  geometric  properties  as  length  and  curvature. 
This  can  be  done  by  defining  three  mutually  perpendicular  unit  vectors  that  form 
the  so-called  moving  frame  specially  adapted  to  a  path  x.  Our  study  takes  us 
briefly  into  the  branch  of  mathematics  called  differential  geometry,  an  area  where 
calculus  and  analysis  are  used  to  understand  the  geometry  of  curves,  surfaces, 
and  certain  higher-dimensional  objects  (called  manifolds). 


3.2  |  Arclength  and  Differential  Geometry 


Length  of  a  Path  

For  now,  let  x:  [a,  b]  —>  R3  be  a  C1  path  in  R3.  Then  we  can  approximate  the 
length  L  of  x  as  follows:  First,  partition  the  interval  [a,  b]  into  n  subintervals. 
That  is,  choose  numbers  to,  t\,  ...,/„  such  that  a  =  to  <  t\  <  ■  ■■  <  t„  =  b.  If, 
for  i  =  1, . . . ,  n,  we  let  As,  denote  the  distance  between  the  points  x(r,_i)  and 
x(ti)  on  the  path,  then 

n 

1=1 

(See  Figure  3.16.)  We  have  x(f)  =  (x(t),  y(t),  z(t)),  so  that  the  distance  formula 
(i.e.,  the  Pythagorean  theorem)  implies 

Ast  =  J  Ax2  +  Ay,2  +  Az2, 

where  Axj  =  x(t{)  -  x(ti-\),  Ayt  =  y(u)  -  y(ti-\),  and  Az,-  =  z{U)  -  zft-i).  It 
is  entirely  reasonable  to  hope  that  the  approximation  in  (1)  improves  as  the  Att 's 
become  closer  to  zero.  Hence,  we  define  the  length  L  of  x  to  be 

n 

L=     lim     Y^AXi2  +  Ay,-2  +  Az,-2.  (2) 

max  Ar,-^0  ~—f 
i=l 

Now,  we  find  a  way  to  rewrite  equation  (2)  as  an  integral.  On  each  subinterval 
fj],  apply  the  mean  value  theorem  (three  times)  to  conclude  the  following: 

1.  There  must  be  some  number  t*  in  [£,-_i,  ?,  ]  such  that 

x(tt)  -  x(»,_i)  =  x'(ff)fe  -  fj_i); 

that  is,  Ax,  =  x'{t*)Att. 

2.  There  must  be  another  number  t**  in  [/,  i ,  r,]  such  that 

Ayi=y'(t**)Att. 

3.  There  must  be  a  third  number  t***  in  [/,_!,  f,  ]  such  that 

Az,  =z'(t***)Ath 
Therefore,  with  a  little  algebra,  equation  (2)  becomes 

L=     lim     V  Jx'{tff  +  y'(t**)2  +  z'(t***)2  Ath  (3) 

max  Ar,^0  z — ■f  V 
(  =  1 

When  the  limit  appearing  in  equation  (3)  is  finite,  it  gives  the  value  of  the  definite 
integral 

b 

Jx>(t)2  +  y>(t)2  +  z'(t)2dt. 

Note  that  the  integrand  is  precisely  ||x'(f)||,  the  speed  of  the  path.  (This  makes 
perfect  sense,  of  course.  Speed  measures  the  rate  of  distance  traveled  per  unit 
time,  so  integrating  the  speed  over  the  elapsed  time  interval  should  give  the  total 
distance  traveled.)  Moreover,  it's  not  hard  to  see  how  we  should  go  about  defining 
the  length  of  a  path  in  R"  for  arbitrary  n. 


L 


204       Chapter  3  |  Vector- Valued  Functions 


DEFINITION  2.1  The  length  L(x)  of  a  C1  path  x:  [a,  b]  ->  R"  is  found  by 
integrating  its  speed: 


=  f  llx' 

J  a 


L(x)  =  /  \\x'(t)\\dt. 


Figure  3.17  AC1  path. 


x(a) 


Figure  3.1 

x: [a,  b] 


x(b) 


8  Apiecewise  C1  path 
R3. 


EXAMPLE  1  To  check  our  definition  in  a  well-known  situation,  we  compute 
the  length  of  the  path 

x:  [0,  2n]  — >  R2,     x(/)  =  (a  cost,  a  sin t),     a  >  0. 


We  have 


so 


x'(r)  =  — a  sinr  i  +  a  cos?  j, 


||x'(f)ll  =  V o2  sin2 1  +  a2  cos2  f  =  a. 
Thus,  Definition  2.1  gives 


L(x) 


2,t 


:  dt  =  lit  a. 


Since  the  path  traces  a  circle  of  radius  a  once,  the  length  integral  works  out  to  be 
the  circumference  of  the  circle,  as  it  should.  ♦ 

EXAMPLE  2   For  the  helix  x(t)  =  (a  cost,  a  sint,  bt),  0  <  t  <  2ir,  we  have 

x'(t)  =  —a  sin  t  i  +  a  cos  t  j  +  b  k, 
so  that  ||x'(0||  =  Va2  +  £2,  and 


L(x) 


=  /     V<s2  +  b2dt  =  +  b2. 

Jo 


When  b  =  0,  the  helix  reverts  to  a  circle  and  the  length  integral  agrees  with  the 
previous  example.  ♦ 

Although  we  have  defined  the  length  integral  only  for  C1  (or  "smooth- 
looking")  paths,  there  is  no  problem  with  extending  our  definition  to  the  piecewise 
C1  case.  By  definition,  a  C1  path  is  one  with  a  continuously  varying  velocity  vec- 
tor, and  so  it  typically  looks  like  the  path  in  Figure  3. 1 7.  A  piecewise  C 1  path  is  one 
that  may  not  be  C1  but  instead  consists  of  finitely  many  C1  chunks.  A  continuous, 
piecewise  C1  path  that  is  notC1  typically  looks  like  the  path  in  Figure  3. 18.  Each 
of  the  three  portions  of  the  path  defined  for  (i)  a  <  t  <  t\,  (ii)  t\  <  t  <  h,  and 
(iii)  t2  <  t  <  b  is  of  class  C 1 ,  but  the  velocity,  if  nonzero,  would  be  discontinuous 
at  t  =  t\  and  t  =  t2  -  To  define  the  length  of  a  piecewise  C1  path,  all  we  need  do  is 
break  up  the  path  into  its  C1  pieces,  calculate  the  length  of  each  piece,  and  add  to 
get  the  total  length.  For  the  piecewise  C 1  path  shown  in  Figure  3.18,  this  means  we 
would  take 

rt\  pt2  rb 

/    \\x'(t)\\dt+      \\x'(t)\\dt+  ||x'(f)||df 

J  a  J  t\  J  t2 

to  be  the  length. 


3.2  |  Arclength  and  Differential  Geometry 


Warning  Even  if  a  path  is  continuous,  the  definite  integral  in  Definition  2.1 
may  fail  to  exist.  An  example  of  such  an  unfortunate  situation  is  furnished  by  the 
pathx:  [0,  1]  R2, 

1 

t  sin 

x(t)  =  (t,  y(t)),     where  y(t) 


t  sin  -    if  t  0 


0 


iff  =  0 


Such  a  path  is  called  nonrectifiable.  It  is  a  fact  that  any  C1  path  with  endpoints 
is  rectifiable,  which  is  why  we  made  such  a  condition  part  of  Definition  2.1. 


The  Arclength  Parameter 

The  calculation  of  the  length  of  a  path  is  not  only  useful  (and  moderately  inter- 
esting) in  itself,  but  it  also  provides  a  way  for  us  to  reparametrize  the  path  with 
a  parameter  that  depends  solely  on  the  geometry  of  the  curve  traced  by  the  path, 
not  on  the  way  in  which  the  curve  is  traced. 

Let  x  be  any  C1  path  and  assume  that  the  velocity  x'  is  never  zero.  Fix  a  point 
P0  on  the  path  and  let  a  be  such  that  x{a)  =  Pq.  We  define  a  one-variable  function 
s  of  the  given  parameter  t  that  measures  the  length  of  the  path  from  Pq  to  any 
other  (moving)  point  P  by 


Figure  3.19  The  arclength 
reparametrization. 


(See  Figure  3.19.  The  Greek  letter  tau,  t,  is  used  purely  as  a  dummy  variable — 
the  standard  convention  is  never  to  have  the  same  variable  appearing  in  both  the 
integrand  and  either  of  the  limits  of  integration.)  If  t  happens  to  be  less  than  a, 
then  the  value  of  s  in  formula  (4)  will  be  negative.  This  is  nothing  more  than  a 
consequence  of  how  the  "base  point"  Po  is  chosen. 

Here's  how  to  get  the  new  parameter:  From  formula  (4)  and  from  the  funda- 
mental theorem  of  calculus, 


ds 
dt 


d 
dt 


J  a 


x\x)\\dx  =  \\x'(t)\\  =  speed. 


(5) 


Since  we  have  assumed  that  x'(t)  ^  0,  it  follows  that  ds/dt  is  nonzero.  Hence, 
ds/dt  is  always  positive,  so  s  is  a  strictly  increasing  function  of  t.  Thus,  s  is, 
in  fact,  an  invertible  function;  that  is,  it  is  at  least  theoretically  possible  to  solve 
the  equation  s  =  s(t)  for  t  in  terms  of  s.  If  we  imagine  doing  this,  then  we  can 
reparametrize  the  path  x,  using  the  arclength  parameter  s  as  independent  variable. 

EXAMPLE  3  For  the  helix  x(t )  =  (a  cost,  a  sin  t,  bt),  if  we  choose  the  "base 
point"  P0  to  be  x(0)  =  (a,  0,  0),  then  we  have 

s(t)  =  f  \\x\r)\\dr=  f  ja2  +  b2dx  =  ja2  +  b2t, 
Jo  Jo 

so  that 

s  =  y/a2  +  b2t, 


Chapter  3  |  Vector-Valued  Functions 


or 

s 

Va2  +  b2 ' 

(What  the  preceding  tells  us  is  that  this  reparametrization  just  rescales  the  time 
variable.)  Hence,  we  can  rewrite  the  helical  path  as 

/        /      s      \  (      s       \  bs 

x(s)  =  la  cos      .  ,  a  sin      .  ,  . 

V        \s/a2  +  b2)  \y/a2  +  b2)    Va2  +  b2 

EXAMPLE  4  The  explicit  determination  of  the  arclength  parameter  for  a  given 
parametrized  path  is  a  delicate  matter.  Consider  the  path 

Thenx'(0  =  (1,  V2t,  t2)  and,  if  we  take  the  base  point  to  be  x(0)  =  (0,  0,  0),then 

s(t)=  f  y/l  +  2x2  +  x4dx 
Jo 

f'  f'  ?3 

=  /  v/(1  +  t2)2Jt  =  /  (l  +  x2)dx  =  t  +  -. 
Jo  Jo  3 

On  the  other  hand  the  path  y(t)  =  (?,  f2,  f3)  is  quite  similar  to  x,  yet  it  has 
no  readily  calculable  arclength  parameter.  In  this  case,  y'(0  =  (1,  2f,  3?2)  and  the 
resulting  integral  for  s(t)  is 


s(t)=  f  y/\+Ax2  +  9xAdx. 
Jo 


It  can  be  shown  that  this  integral  has  no  "closed  form"  formula  (i.e.,  a  formula 
that  involves  only  finitely  many  algebraic  and  transcendental  functions).  ♦ 

The  significance  of  the  arclength  parameter  s  is  that  it  is  an  intrinsic  param- 
eter; it  depends  only  on  how  the  curve  itself  bends,  not  on  how  fast  (or  slowly) 
the  curve  is  traced.  To  see  more  precisely  what  this  means,  we  resort  to  the  chain 
rule.  Consider  s  as  an  intermediate  variable  and  t  as  a  final  variable.  Then  we 
have 

ds 

x(t)  =  x  (s) —  by  the  chain  rule, 

dt 

=  As)\\x'(t)\\       by  (5). 
Since  x'(t)  ^  0,  we  can  solve  for  x'(s)  to  find 

x'(s)= — —.  (6) 
l|x'(OII 

Therefore,  x'(s)  is  precisely  the  normalization  of  the  original  velocity  vector,  and 
so  it  is  a  unit  vector.  Hence,  the  reparametrized  path  x(s )  has  unit  speed,  regardless 
of  the  speed  of  the  original  path  x(f).  (This  result  makes  good  geometric  sense, 
too.  If  arclength,  rather  than  time,  is  the  parameter,  then  speed  is  measured  in 
units  of  "length  per  length,"  which  necessarily  must  be  one.) 

The  only  unfortunate  note  to  our  story  is  that  the  integral  in  formula  (4)  is 
usually  impossible  to  compute  exactly,  thus  making  it  impossible  to  compute  s 
as  a  simple  function  of  t.  (The  case  of  the  helix  is  a  convenient  and  rather  special 


3.2  |  Arclength  and  Differential  Geometry 


exception.)  One  generally  prefers  to  work  indirectly,  letting  the  chain  rule  come 
to  the  rescue.  We  shall  see  this  indirect  approach  next. 


The  Unit  Tangent  Vector  and  Curvature  

Let  x:  /  c  R  — ■>•  R3  be  a  C3  path  and  assume  that  x'  is  never  zero. 


DEFINITION  2.2  The  unit  tangent  vector  T  of  the  path  x  is  the  normal- 
ization of  the  velocity  vector;  that  is, 

v  x'(0 


T  = 


Ml  l|x'(OII 


We  see  from  Definition  2.2  that  the  unit  tangent  vector  is  undefined  when  the 
speed  of  the  path  is  zero.  Also  note  that,  from  equation  (6),  T  is  dx/ds,  where  s 
is  the  arclength  parameter.  Geometrically,  T  is  the  tangent  vector  of  unit  length 
that  points  in  the  direction  of  increasing  arclength,  as  suggested  by  Figure  3.20. 

EXAMPLE  5    For  the  helix  x(f)  =  (a  cost,  a  sin  t,  bt),  we  have 

x'(f)       —  a  sin?  i  +  a  cos  t  j  +  bk. 

T(f)  =  — —  =    J  . 

I|x'(0ll  Ja2  +  b2 

On  the  other  hand  if  we  parametrize  the  helix  using  arclength  so  that 

bs 


then 


x(s)  =    a  cos      ,  ] ,  a  sin  . 

\Vfl2  +  W  \y/a2  +  b2)  Va^Tb2 


T(s)  =  x'O)  =    .  =     =  sin  i  +  cos 


■s/a2  +  b2      V  V«2  +  b2  )       y/a2  +  b2       \-Ja2  +  b2 
b 

+ 


Va2  +  b2 

This  agrees  (as  it  should)  with  the  first  expression  for  T,  since  s  =  -J a2  +  b2  t, 
as  shown  in  Example  3.  ♦ 

Using  the  unit  tangent  vector,  we  can  define  a  quantity  that  measures  how 
much  a  path  bends  as  we  travel  along  it.  To  do  so,  note  the  following  key  facts: 

PROPOSITION  2.3  Assume  that  the  path  x  always  has  nonzero  speed.  Then 

1.  dT/dt  is  perpendicular  to  T  for  all  t  in  /  (the  domain  of  the  path  x). 

2.  \\dT/dt  ||  \t=to  equals  the  angular  rate  of  change  (as  t  increases)  of  the  direc- 
tion of  T  when  t  =  to. 


PROOF  (You  can  omit  reading  this  proof  for  the  moment  if  you  are  interested  in 
the  main  flow  of  ideas.)  To  prove  part  1 ,  we  have 


T(0-T(0  =  1, 


Chapter  3  |  Vector-Valued  Functions 


since  T  is  a  unit  vector.  Hence, 


T(r0  +  At) 


T(t0) 


Figure  3.21  The  vector  triangle 
used  in  the  proof  of 
Proposition  2.3. 


dt 


(T  •  T)  =  0, 


because  the  derivative  of  a  constant  is  zero.  Also  we  have 

d                  dT  dT 
—  (T-T)  =  T  1  T, 

dt  dt  dt 

by  the  product  rule  (Proposition  1.4).  Thus, 

dT 

2T  =  0. 

dt 

Therefore,  T  is  always  perpendicular  to  dT/dt.  (See  Proposition  1.7.) 

Now  we  prove  part  2.  Because  T  is  a  unit  vector  for  all  t,  only  its  direction 
can  change  as  t  increases.  This  angular  rate  of  change  of  T  is  precisely 

lim  — , 

Af->0+  At 

where  AO  comes  from  the  vector  triangle  shown  in  Figure  3.21.  To  make  the 
argument  technically  simpler,  we  shall  assume  that  AT  ^  0.  We  claim  that 

AO 


lim 

Ar^0+  IIATI 


1. 


(7) 


Then,  from  equation  (7), 


AO                AO  ||AT|| 
lim  —  =  lim  

AI-S-0+  At       Af^0+  II  AT 


At 


=  lim 


AO 


lim 


IATII 


a^o+  II  AT  II  Af^o+  At 


=  1  •  lim 


IATII 


Ar^0+  At 

Since  At  is  assumed  to  be  positive  in  the  limit,  we  may  conclude  that 
lim 


AT 

dT 

— -  =  lim 

dt 

At  Ar^0+ 

At 

as  desired. 

To  establish  equation  (7),  the  law  of  cosines  applied  to  the  vector  triangle  in 
Figure  3.21  implies 

||  AT||2  =  ||T(f  +  Af)||2  +  ||T(f)||2  -  2||T(t  +  At)||  ||T(t)||  cos  AO 

=  2  -  2  cos  AO, 

because  T  is  always  a  unit  vector.  Thus, 

AO 


lim 


lim 


AO 


Ar^o+  ||AT||      a^o+  72  -  2  cos  AG 

AO 

=   lim  - 

A'^0+ ^2  •  2(sin2(A6>/2)) 

from  the  half-angle  formula,  and  so 

AO  AO/2 

lim   =   lim   =  1, 

Ar^o+  ||AT||      Ar^o+  sin(A<9/2) 

from  the  well-known  trigonometric  limit  (or  from  L'Hopital's  rule). 


3.2  j  Arclength  and  Differential  Geometry  209 

Part  2  of  Proposition  2.3  provides  a  precise  way  of  measuring  the  bending  of 
a  path. 


DEFINITION  2.4  The  curvature  k  of  a  path  x  in  R3  is  the  angular  rate  of 
change  of  the  direction  of  T  per  unit  change  in  distance  along  the  path. 


The  reason  for  taking  the  rate  of  change  of  T  per  unit  change  in  distance  in  the 
definition  of  k  is  so  that  the  curvature  is  an  intrinsic  quantity  (which  we  certainly 
want  it  to  be).  Figure  3.22  should  help  you  develop  some  intuition  about  k. 


Figure  3.22  In  the  left  figure,  k  is  not  large,  since  the 
path's  unit  tangent  vector  turns  only  a  small  amount  per 
unit  change  in  distance  along  the  path.  In  the  right 
figure,  k  is  much  larger,  because  T  turns  a  great  deal 
relative  to  distance  traveled. 


Because  ||dT/<2r||  measures  the  angular  rate  of  change  of  the  direction  of  T 
per  unit  change  in  parameter  (by  part  2  of  Proposition  2.3)  and  ds/dt  is  the  rate 
of  change  of  distance  per  unit  change  in  parameter,  we  see  that 


where  the  last  equality  holds  by  the  chain  rule.  It  is  formula  (8)  that  we  will  use 
when  making  calculations. 

EXAMPLE  6   For  the  circle  x(?)  =  {a  cos  t,  a  sin  t),  0  <  t  <2tt, 

ds 

x  (?)  =  —  a  sin?  i  +  a  cost  j,        ||x  (?)||  =  —  =  a, 

dt 

so  that 

xYf) 

T(?)  =  =  —  sin? i  +  cos?  j. 

I|x'(?)|| 


Hence, 


\\dT/dt\\      1  .  1 

k  =  =  -||  -cos?i-sm?j||  =  - 

ds/dt        a  a 


Thus,  we  see  that  the  curvature  of  a  circle  is  always  constant  with  value  equal 
to  the  reciprocal  of  the  radius.  Therefore,  the  smaller  the  circle,  the  greater  the 
curvature.  (Draw  a  sketch  to  convince  yourself.)  ♦ 


210 


Chapter  3  |  Vector-Valued  Functions 


EXAMPLE  7   If  a  and  b  are  constant  vectors  in  R3  and  the  path 

x(t)  =  a  t  +  b 


traces  a  line.  We  have 


so 


x'(0  =  a, 
ds 


Hence, 


dt 


T(/)  = 


=  a  . 


which  is  a  constant  vector.  Thus,  T'(f)  =  0  and  formula  (8)  implies  immedi- 
ately that  k  is  zero,  which  agrees  with  the  intuitive  fact  that  a  line  doesn't 
curve.  ♦ 

EXAMPLE  8    Returning  to  our  friend  the  helix 

x(t)  =  (a  cos  t ,  a  sin  t ,  bt), 

we  have  already  seen  that 


ds 
dt 


=  ja2  +  b2    and    T(f)  = 


-a  sin  /  i  +  a  cos  /  j  +  b  k 

Va2  +  b2 


Thus,  formula  (8)  gives 


1 


sja2  +  b2 


-a  cos  /  i  —  a  sin?  j 


Va2  +  b2 


a2  +  b2' 


We  see  that  the  curvature  of  the  helix  is  constant,  just  like  the  circle.  In  fact,  as  b 
approaches  zero,  the  helix  degenerates  to  a  circle,  and  the  resulting  curvature  is 
consistent  with  that  of  Example  6. 

We  can  also  compute  the  curvature  from  the  parametrization  given  by  arc- 
length.  The  same  helix  is  also  described  by 


\(s)  =    a  cos 


Vfl2  +  b2 


,  a  sin 


bs 


vV  +  W  Va2  +  W 


and  we  have 
dx 


T(s)  = 


ds 


sin 


+ 


Va2  +  b2      \Ja2  +  b2 
b 


+ 


cos 


Vfl2  +  b2       \y/a2  +  b2 


■s/a2  +  b2 
We  can,  therefore,  compute 
dT  a 
ds 


cos 


a2  +  b2       \Ja2  +  b2 
and  hence,  from  formula  (8),  that 

dT 


sin 


+  b2       \s/a2  +  b2 


h 


ds 


a2  +  b2 


which  checks. 


3.2  |  Arclength  and  Differential  Geometry 


The  Moving  Frame  and  Torsion  

We  now  introduce  a  triple  of  mutually  orthogonal  unit  vectors  that  "travel"  with  a 
given  path  x:  /  —>  R3,  known  as  the  moving  frame  of  the  path.  (Note:  In  general, 
the  term  "frame"  means  an  ordered  collection  of  mutually  orthogonal  unit  vectors 
in  R".)  These  vectors  should  be  thought  of  as  a  set  of  special  vector  "coordinate 
axes"  that  move  from  point  to  point  along  the  path. 

To  begin,  assume  that  (i)  x'(0  #  0  and  (ii)  x'(t)  x  x"(f)  ^  0  for  all  t  in  /. 
(The  first  condition  assures  us  that  x  never  has  zero  speed  and  the  second  that  x 
is  not  a  straight-line  path.)  Then  the  first  vector  of  the  moving  frame  is  just  the 
unit  tangent  vector: 

dx  =  x'(Q 
ds      ||x'(OII ' 

(Now  you  see  why  condition  (i)  is  needed.)  For  a  second  vector  orthogonal  to  T, 
recall  that  part  1  of  Proposition  2.3  says  that  dT/dt  must  be  perpendicular  to  T. 
Hence,  we  define 


(That  dl/dt  is  not  zero  follows  from  assumptions  (i)  and  (ii).)  The  vector  N  is 
called  the  principal  normal  vector  of  x.  By  the  chain  rule,  N  is  also  given  by 

dT/ds 

N  =  .  (10) 

\\dT/ds\\ 

Since  k  =  \\dT/ds\\  by  formula  (8),  we  also  see  that 


At  a  given  point  P  along  the  path,  the  vectors  T  and  N  (and  also  the  vectors 
x'  and  x")  determine  what  is  called  the  osculating  plane  of  the  path  at  P.  (See 
Figure  3.23.)  This  is  the  plane  that  "instantaneously"  contains  the  path  at  P.  (More 


Osculating  plane 


Figure  3.23  The  osculating  plane  of  the  path  x  at  the 
point  P. 


212       Chapter  3  |  Vector- Valued  Functions 


precisely,  it  is  the  plane  obtained  by  taking  points  P\  and  P2  on  the  path  near  P 
and  finding  the  limiting  position  of  the  plane  through  P,  Pi,  and  P%  as  Pi  and 
P2  approach  P  along  x.  The  word  "osculating"  derives  from  the  Latin  osculare, 
meaning  "to  kiss.") 

Now  that  we  have  defined  two  orthogonal  unit  vectors  T  and  N,  we  can 
produce  a  third  unit  vector  perpendicular  to  both: 


B  =  T  x  N. 


(12) 


The  vector  B,  called  the  binormal  vector,  is  defined  so  that  the  ordered  triple 
(T,  N,  B)  is  a  right-handed  system.  Thus,  B  is  a  unit  vector  since 


IBM  =  II T II 


sin  -  =  1  ■  1  •  1  =  1. 

2 


EXAMPLE  9  For  the  helix  x(/)  =  (a  cos  t,  a  sin  t,  bt),  the  moving  frame  vec- 
tors are 


T(f)  = 


-a  sin  t  i  +  a  cos  t  j  +  b  k 

■J  a1  +  b2 


(as  we  have  already  seen), 


T'(f)       (—a  cos  t  i  —  a  sinf  j)/V«2  +  b2 
fs(t)  =  — —  =  ^=^=  =  —  cos  1 1  —  sin  t  j , 


|T'(t)ll 


a/Va2  +  b2 


and 


B(0  =  T  x  N  = 


i  j  k 

-a  sinf/Va2  +  bz      a  cos?/V«2  +  b2      b/-Ja2  +  b2 
—  cos  t  —  sin  t  0 


sinf   i-      ,  ,     Fi  cosf  j+      /  i  k- 


\«Ja2  +  b2       J       \V<32  +  b2        J       Wfl2  +  b2 

Equation  (11)  says  that  the  derivative  of  T  (with  respect  to  arclength)  is  a 
scalar  function  (namely,  the  curvature)  multiple  of  the  principal  normal  N.  This 
is  not  surprising,  since  N  is  defined  to  be  parallel  to  the  derivative  of  T.  A  more 
remarkable  result  (see  the  addendum  at  the  end  of  this  section)  is  that  the  derivative 
of  the  binormal  vector  is  also  always  parallel  to  the  principal  normal;  that  is, 

dB 

—  =  (scalar  function)  N. 

ds 

The  standard  convention  is  to  write  this  scalar  function  with  a  negative  sign,  so 
we  have 


dB 

—  =  -tN.  (13) 
ds 


The  scalar  function  r  thus  defined  is  called  the  torsion  of  the  path  x.  Roughly 
speaking,  the  torsion  measures  how  much  the  path  twists  out  of  the  plane,  how 


3.2  |  Arclength  and  Differential  Geometry 


"three-dimensional"  x  is.  Note  that,  according  to  our  conventions,  the  curvature 
k  is  always  nonnegative  (why?),  while  r  can  be  positive,  negative,  or  zero. 

EXAMPLE  10  Consider  again  the  case  of  circular  motion.  Thus,  let  x(/)  = 
(a  cost,  a  sin /).  Then,  as  shown  in  Example  6, 


T(0 


x'(t) 


|x'(OII 


sin?  i  + cos?  j,    and  k 


dT 


ds 


Now  we  calculate 


T'(0 

N  =  =  —  cos  M-sinn', 

l|T'(0ll 

B  =  TxN  =  k,    a  constant  vector. 

Hence,  dB/ds  =  0,  so  there  is  no  torsion.  This  makes  sense,  since  a  circle  does 
not  twist  out  of  the  plane.  ♦ 

EXAMPLE  1 1    Let  x(/)  =  (e'  cos  /,  e'  sin  / ,  e').  We  calculate  T,  N,  and  B  and 
identify  the  curvature  and  torsion  of  x. 
To  begin,  we  have 

x'(t)       e'(cos/  —  sin/)i  +  e'(cos/  +  sin/)j  +  e'  k 

||x'(0||  =  Vfe< 


1 

71 


((cos  t  —  sin  t)  i  +  (cos  t  +  sin  t )  j  +  k) . 


From  this,  we  may  compute 

dT  dT/dt  ^(-(sin? +  cos0i  +  (cosf  -  sinf)j) 
ds      ds/dt  s/3e' 

e-' 

=  — (— (sinf  +  cosf)i  +  (cos  t  —  sinf)j), 


so  that  the  curvature  is 


dT 


ds 


sflt 


3 


Now  we  determine  the  remainder  of  the  moving  frame: 
T'(0  1 

N  =  -t:  jr  =  —=(—  (sinf  +  cos  t)i  +  (cosf  —  sin/)  j), 

T'(0  V2 


B  =  T  x  N  =  —((sinf  -  cos/)i  -  (sin/  +  cos/)j  +  2k). 
V  6 

Finally,  to  find  the  torsion,  we  calculate 

dB  _  dB/dt  _  ^((cos/  +  sin/)i  +  (sin/ -cos/)j) 
ds      ds/dt  yfZe* 

((cos  f  +  sin  /)  i  +  (sin  /  —  cos  /)  j) 


3^2 


-N, 


so 


T  = 


214       Chapter  3  |  Vector-Valued  Functions 


Figure  3.24  Any  vector  in  the 
plane  perpendicular  to  T  can  be 
used  for  N. 


EXAMPLE  12  If  a  and  b  are  vectors  in  R3,  then  the  straight-line  path  x(t)  = 
at  +  bhas,  as  we  saw  in  Example  7,  T  =  a/||a||.  Thus,  both  dT/df  anddT/ds  are 
identically  zero.  Hence,  k  =  0  (as  shown  in  Example  7)  and  N  cannot  be  defined 
using  formula  (9).  From  geometric  considerations,  any  unit  vector  perpendicular 
to  T  can,  in  principle,  be  used  for  N.  (See  Figure  3.24.)  If  we  choose  one  such 
vector,  then  B  can  be  calculated  from  formula  (12).  Since  T,  N,  and  B  are  all 
constant,  r  must  be  zero.  This  is  an  example  of  a  moving  frame  that  is  not 
uniquely  determined  by  the  path  x  and  serves  to  illustrate  why  the  assumption 
x'  x  x"  /  0  was  made.  ♦ 

It  is  important  to  realize  that  the  moving  frame,  curvature,  and  torsion  are 
quantities  that  are  intrinsic  to  the  curve  traced  by  the  path.  That  is,  any  parame- 
trized path  that  traces  the  same  curve  (in  the  same  direction)  must  necessarily 
have  the  same  T,  N,  B  vector  functions  and  the  same  curvature  and  torsion.  This 
is  because  all  of  these  quantities  can  be  denned  entirely  in  terms  of  the  intrinsic 
arclength  parameter  s.  (See  Definition  2.2  and  formulas  (6),  (8),  (10),  (1 1),  (12), 
and  (13).) 

Another  important  fact  is  that  the  curvature  function  k  and  the  torsion  function 
t  together  determine  all  the  geometric  information  regarding  the  shape  of  the 
curve,  except  for  the  curve's  particular  position  in  space.  To  be  more  precise,  we 
have  the  following  theorem,  whose  proof  we  omit: 


THEOREM  2.5  Let  s  be  the  arclength  parameter  and  suppose  C\  and  C2  are 
two  curves  of  class  C3  in  R3 .  Assume  that  the  corresponding  curvature  functions 
k\  and  K2  are  strictly  positive.  Then  if  k\(s)  =  K2(s)  and  t\(s)  =  t2(s),  the  two 
curves  must  be  congruent  (in  the  sense  of  high  school  geometry).  In  fact,  given 
any  two  continuous  functions  k  and  r,  where  k{s)  >  0  for  all  s  in  the  closed 
interval  [0,  L],  there  is  a  unique  curve  parametrized  by  arclength  on  [0,  L]  (up  to 
position  in  space)  whose  curvature  and  torsion  are  k  and  r,  respectively. 


Tangential  and  Normal  Components  of  Velocity  and  Accel- 
eration; Other  Curvature  Formulas  — 

As  we  have  seen,  the  moving  frame  provides  us  with  an  intrinsic  set  of  vectors, 
like  coordinate  axes,  that  are  special  to  the  particular  curve  traced  by  a  path.  In 
contrast,  the  velocity  and  acceleration  vectors  of  a  path  are  definitely  not  intrinsic 
quantities  but  depend  on  the  particular  parametrization  chosen  as  well  as  on  the 
shape  of  the  path.  (The  speed  of  a  path  is  entirely  independent  of  the  geometry 
of  the  curve  traced.)  We  can  get  some  feeling  for  the  relationship  between  the 
intrinsic  notion  of  the  moving  frame  and  the  extrinsic  quantities  of  velocity  and 
acceleration  by  expressing  the  latter  two  vector  functions  in  terms  of  the  moving 
frame  vectors. 

Thus,  we  begin  with  a  C2  path  x:  /  — >  R3  having  x'  7^  0  and  x'  x  x"  ^  0. 
For  notational  convenience,  let  s  denote  ds/dt  and  s  denote  d2s/dt2.  From 
Definition  2.2,  we  know  that  T  =  v/||v||  and  so,  since  the  speed  s  =  ds/dt  =  ||v||, 
we  have 


v(f)  =  sT.  (14) 


3.2  |  Arclength  and  Differential  Geometry 


Figure  3.25  Decomposition  of 
acceleration  a  into  tangential  and 
normal  components. 


This  formula  says  that  the  velocity  is  always  parallel  to  the  unit  tangent  vector, 
something  we  know  well.  To  obtain  a  similar  result  for  acceleration,  we  can 
differentiate  (14)  and  apply  the  product  rule: 


a(0  =  V(t) 


dt 


(sT)  =  ST  +  s 


dT 
dt  ' 


(15) 


Next,  we  express  dT/dt  in  terms  of  the  T,  N,  B  frame.  Formula  (11)  gives 
the  derivative  of  dT/ds  in  terms  of  N.  The  chain  rule  says  that  dT/ds  = 
(dT/dt)/(ds/dt).  Thus,  from  formula  (1 1),  we  have 

dT  dT 

——  =  s—  =  skN. 

dt  ds 

Hence,  we  may  rewrite  equation  (15)  as 


a(r)  =  j'T  +  k  j2N. 


(16) 


Warning  s  =  d2s/ dt2  is  the  derivative  of  the  speed  which  is  a  scalar  function. 
The  acceleration  a  is  the  derivative  of  velocity  and  so  is  a  vector  function. 

Note  that  formula  (16)  shows  that  the  acceleration  has  no  component  in  the 
direction  of  the  binormal  vector  B.  Therefore,  both  velocity  and  acceleration  are 
vectors  that  lie  in  the  osculating  plane  of  the  path.  (See  Figure  3.25.) 

At  first  glance,  it  may  not  appear  to  be  especially  easy  to  use  formula  (16) 
to  resolve  acceleration  into  its  tangential  and  normal  components  because  of  the 
curvature  term.  However, 

||a||2  =  a  •  a  =  (j'T  +  k j2N)  •  (ST  +  /fj2N)  =  s2  +  (ks2)2, 

since  T  and  N  are  perpendicular  vectors.  Consequently,  we  may  calculate  the 
components  as  follows: 


Tangential  component  of  acceleration  =  atang  =  s. 


Normal  component  of  acceleration  =  an0rm  =  ks2  =  J  ||a||z  —  a 


'tang- 


EXAMPLE  1 3   Let  x(Q  =  (t ,  It ,  t2).  Then  v(/)  =  i  +  2j  +  2/k  and  a(f )  =  2k. 

We  have  s  =  ||v(f)ll  =  V5  +  4f2.  Therefore, 


At 


V5  +  At2 ' 
Since  ||a||  =  2,  we  see  that 


I  a  II 2  -  fitting 


16/2  2^5 


5  +  4f2      V5  +  At 2 ' 


Formulas  (14)  and  (16)  enable  us  to  find  an  alternative  equation  for  the 
curvature  of  the  path.  We  simply  calculate  that 

v  x  a  =  (iT)  x  (sT  +  /ci2N)  =  ss(T  x  T)  +  kP(T  x  N)  =  ks3B. 


216       Chapter  3  I  Vector- Valued  Functions 

Recalling  that  s  =  ||v||,  we  have,  by  taking  magnitudes, 
||v  x  a||  =  k||v||3||B||  =K ||v||3, 
since  B  is  a  unit  vector.  Thus, 


||vxa|| 

(17) 

K  =   z—. 

IMP 

This  relatively  simple  formula  expresses  the  curvature  (an  intrinsic  quantity) 
in  terms  of  the  nonintrinsic  quantities  of  velocity  and  acceleration. 

EXAMPLE  14   For  the  path  \(t)  =  (2t3  +  l,t4,  t5),  we  have 


and 


You  can  check  that 


v(0  =  6th  +  4f3j  +  5r4k 


a(0  =  12ri  +  I2t2\  +  20f3k. 


|v||  =  t2^25t4  +  I6t2  +  36 


and 


||v  x  a||  =  ||4f4(5f2i  -  I5tj  +  6k)||  =  4fV25f4  +  225/2  +  36. 

Therefore,  formula  (17)  yields 

|| v  x  a||  _  4(25f4  +  225r2  +  36)1/2 
*  ~    ||v||3    ~  t2(25t4  +  16f2  +  36)3/2  ' 

which  is  certainly  a  more  convenient  way  to  determine  curvature  in  this  case.  ♦ 


Summary   

You  have  seen  many  formulas  in  this  section,  and,  at  first,  it  may  seem  difficult 
to  sort  out  the  primary  statements  from  the  secondary  results.  We  list  the  more 
fundamental  facts  here: 

For  a  path  x:  /  — >•  R3 : 


Nonintrinsic  quantities: 

Velocity  v(0  =  x'(t). 

Speed  %  =  ||v(/)||. 
dt 

Acceleration  a(t)  =  x"(?). 


3.2  |  Arclength  and  Differential  Geometry 


Arclength  function:  (See  Figure  3.26.) 

s(t)  =  /    ||x'(t)||Jt    (basepoint  is  Pq  =  x(a)) 

J  a 


Intrinsic  quantities: 

The  moving  frame: 


Unit  tangent  vector  T  = 

Principal  normal  vector  N  = 
Binormal  vector  B  = 

Curvature  k  = 
dB 


Torsion  r  is  denned  so  that 


ds 


dx 

X 

ds~  ~ 

II* 

dT/ds 

\\dT/ds\\ 

TxN. 

dT 

ds 

-tN 

At) 


dT/dt 
=  \\dT/dt\ 

WdT/dtW 
ds/dt 


Additional  formulas: 

v(?)  =  s  T    (s  is  speed). 

a(t)  =  ST  +  ks2  N    (s  is  derivative  of  speed). 
Ilv  x  all 


Addendum:  More  About  Torsion  and  the 
Frenet-Serret  Formulas 

We  now  derive  formula  (13),  the  basis  for  the  definition  of  the  torsion  of  a  curve. 
That  is,  we  show  that  the  derivative  of  the  binormal  vector  B  (with  respect  to 
arclength)  is  always  parallel  to  the  principal  normal  N  (i.e.,  that  dB/ds  is  a 
scalar  function  times  N).  The  two  main  ingredients  in  our  derivation  are  part  1  of 
Proposition  2.3  and  the  product  rule. 

We  begin  by  noting  that,  since  the  ordered  triple  of  vectors  (T,  N,  B)  forms  a 
frame  for  R3,  any  moving  vector,  including  dB/ds,  can  be  expressed  as  a  linear 
combination  of  these  vectors;  that  is,  we  must  have 


dB 

ds 


a(s)T  +  Z?(s)N  +  c(«)B, 


(18) 


where  a,  b,  and  c  are  appropriate  scalar- valued  functions.  (Because  T,  N,  and 
B  are  mutually  perpendicular  unit  vectors,  any  (moving)  vector  w  in  R3  can  be 
decomposed  into  its  components  with  respect  to  T,  N,  and  B  in  much  the  same 
way  that  it  can  be  decomposed  into  i,  j,  and  k  components — see  Figure  3.27.)  To 
find  the  particular  values  of  the  component  functions  a,  b,  and  c,  it  turns  out  that 


218       Chapter  3  |  Vector-Valued  Functions 


we  can  solve  for  each  function  by  applying  appropriate  dot  products  to  equation 
(18).  Specifically, 

dB 

 T  =  a(s)T  •  T  +  b(s)N  •  T  +  c(s)B  ■  T 

ds 

=  a(s)  ■  1  +  b(s)  •  0  +  c(s)  •  0 
=  a(s), 

and,  similarly, 

dB  dB 

—  -N  =  —.B  =  c(s). 

ds  as 

From  Proposition  1.7,  dB/ds  is  perpendicular  to  B  and,  hence,  c  must  be  zero. 
To  find  a,  we  use  an  ingenious  trick  with  the  product  rule:  Because  T  •  B  =  0,  it 
follows  that  d /ds(T  •  B)  =  0.  Now,  by  the  product  rule, 

d                   dB  dT 
— (TB)  =  T  1  B. 

ds  ds  ds 

Consequently,  (dB/ds)  •  T  =  —(dT/ds)  •  B.  Thus, 

,  .      dB  ^        dT  „ 

a(s)  =  T  =  B 

ds  ds 

=  -*rN-B       by  formula  (11), 
=  0, 


and  equation  ( 1 8)  reduces  to 


dB 

=  b(sys. 

ds 


No  further  reductions  are  possible,  and  we  have  proved  that  the  derivative  of  B  is 
parallel  to  N.  The  torsion  r  can,  therefore,  be  defined  by  r(s)  =  —  b(s). 

Formulas  (1 1)  and  (13)  gave  us  intrinsic  expressions  for  dT/ds  and  dB/ds, 
respectively.  We  can  complete  the  set  by  finding  an  expression  for  dN/ds.  The 
method  is  the  same  as  the  one  just  used.  Begin  by  writing 

JN 

—  =  o(j)T  +  MsyS  +  c(s)B,  (19) 
ds 

where  a,  b,  and  c  are  suitable  scalar  functions.  Taking  the  dot  product  of  equation 
(19)  with,  in  turn,  T,  N,  and  B,  yields  the  following: 

JN 

a(s)  =  —  •  T,       b(s)  =  —  •  N,       c(s)  =  —  •  B. 
ds  ds  ds 

The  "product  rule  trick"  used  here  then  reveals  that 

i/N  „  dT 

a(s)=  —  .T  =  -N.— 
ds  ds 

=  — N-/cN      by  formula  (11) 


and 


JN  JB 
c(s)=        B  =  -N.— 
ds 


=  — N  •  (-tN)       by  formula  (13) 

=  T. 


3.2  |  Exercises  219 


Moreover,  we  may  differentiate  the  equation  N  •  N  =  1  to  find 

b(s) 


dN  dN 
■  N  =  — N  • 


=  -a:T  +  tB. 


ds  ds  ' 

which  implies  that  b(s)  is  zero.  Hence,  equation  (19)  becomes 

dN 

ds 

The  formulas  for  dT/ds,  dN/ds,  and  dB/ds  are  usually  taken  together  as 

T'O)  =  kN 
N'(i)  =  -kT  +  tB 
B'O)  =  -tN 

and  are  known  as  the  Frenet-Serret  formulas  for  a  curve  in  space.  They  are  so 
named  for  Frederic-Jean  Frenet  and  Joseph  Alfred  Serret,  who  published  them 
separately  in  1852  and  1851,  respectively  The  Frenet-Serret  formulas  give  a 
system  of  differential  equations  for  a  curve  and  are  key  to  proving  a  result  like 
Theorem  2.5.  They  are  often  written  in  matrix  form,  in  which  case,  they  have  an 
especially  appealing  appearance,  namely, 


r " 

0 

K 

0  " 

~  T 

N' 

—  K 

0 

X 

N 

B' 

0 

—  T 

0 

B 

3.2  Exercises 


Calculate  the  length  of  each  of  the  paths  given  in  Exercises 
1-6. 


1. 

x(f)  = 

(2t+  l,7-3*),-l  <r  <2 

2. 

x(f)  = 

f2i+  |(2f  +  l)3/2j,  0  <  f  <  4 

3. 

x(f)  = 

(cos  3?,  sin3f,  2?3/2),  0  <  t  <  2 

4. 

x(f)  = 

7i  + t]  +  t2k,  1  <  t  <  3 

5. 

x(f)  = 

(r3,3r2,6f), -1  <r  <2 

6. 

x(f)  = 

(In (cos 0,  cost,  sin f),  f  <  ?  <  f 

7. 

x(f)  = 

Qnt,  t2/2,  y/2t),  1  <  t  <  4 

8. 

x(f)  = 

(2t  cosf,2f  sinr,2V2f2),  0  <  ?  <  3 

9. 

The  path  x(f )  =  (a  cos3 1,  a  sin3  f),  where  a 

tive  constant,  traces  a  curve  known  as  an  astroid  or  a 
hypocycloid  of  four  cusps.  Sketch  this  curve  and  find 
its  total  length.  (Be  careful  when  you  do  this.) 

10.  If  /  is  a  continuously  differentiable  function,  show 
how  Definition  2.1  may  be  used  to  establish  the 
formula 

L  =   /  y/\+{f'(X)Ydx 
J  a 


forthe  length  ofthe  curve  y  =  f(x)  between  (a,  f(a)) 
and  (b,  f(b)). 

1 1 .  Use  Exercise  10  or  Definition  2.1  (or  both)  to  calculate 
the  length  of  the  line  segment  y  =  mx  +  b  between 
(xq,  yo)  and  (xi,  yi).  Explain  your  result  with  an  ap- 
propriate sketch. 

12.  (a)  Calculate  the  length  of  the  line  segment  deter- 

mined by  the  path 

x(t)  =  (alt  +  bua2t  +  b2) 

as  f  varies  from  to  to  t\ . 

(b)  Compare  your  result  with  that  of  Exercise  1 1 . 

(c)  Now  calculate  the  length  of  the  line  segment  deter- 
mined by  the  path  x(f )  =  a  t  +  b  as  t  varies  from 
f0  to  h . 

13.  This  problem  concerns  the  path  x=|f  —  l|i+|r|j, 

-2  <  t  <  2. 

(a)  Sketch  this  path. 

(b)  The  path  fails  to  be  of  class  C1  but  is  piecewise 
Cl .  Explain. 

(c)  Calculate  the  length  of  the  path. 

14.  Consider  the  path  x(f)  =  (e~'  cost,  e~'  sinf). 


Chapter  3  |  Vector-Valued  Functions 


(a)  Argue  that  the  path  spirals  toward  the  origin  as 
t  -»-  +00. 

(b)  Show  that,  for  any  a,  the  improper  integral 


r 

J  a 


\W{t)\\dt 


converges. 

(c)  Interpret  what  the  result  in  part  (b)  says  about  the 
path  x. 

1 5.  Suppose  that  a  curve  is  given  in  polar  coordinates  by 
an  equation  of  the  form  r  =  f(9),  where  /  is  of  class 
C1.  Use  Definition  2.1  to  derive  the  formula 


J  a 


(8)2  +  f(6)2d6 


for  the  length  of  the  curve  between  the  points  (/(a),  a) 
and  (/(/S),  j3)  (given  in  polar  coordinates). 

16.  (a)  Find  the  arclength  parameter  s  =  s(t)  for  the  path 

x(?)  =  eat  cos  bt  i  +  eat  sin  bt  j  +  eat  k. 

(b)  Express  the  original  parameter  /  in  terms  of  s  and, 
thereby,  reparametrize  x  in  terms  of  s. 

Determine  the  moving  frame  {T,  N,  B},  and  compute  the  cur- 
vature and  torsion  for  the  paths  given  in  Exercises  1 7—20. 

17.  x(0  =  5cos3H  +  6t  j  +  5  sin  3?  k 

18.  x(f)  =  (sin?  -  t  cos?)i  +  (cosf  +  t  sinf)j  +  2k, 
t  >  0 

19.  x(r)  =  (t,  \(t  +  I)3/2,  1(1  -  tf'2),  -1  <  t  <  1 

20.  x(f)  =  (e2'  sin/,  e2'  cosf,  1) 

21.  (a)  Use  formula  (17)  in  this  section  to  establish  the 

following  well-known  formula  for  the  curvature 
of  a  plane  curve  y  =  f(x): 


[i  +  (/'(x))2F2' 

(Assume  that  /  is  of  class  C2.) 

(b)  Use  your  result  in  (a)  to  find  the  curvature  of 
y  =  In  (shut). 

22.  (a)  Let  x(s)  =  {x(s),  y(s))  be  a  plane  curve  para- 
metrized by  arclength.  Show  that  the  curvature  is 
given  by  the  formula 

K  =  \x'y" -x"y'\. 


path  x  and,  separately,  plot  the  curvature  k  as  a  function  of  t 
over  the  indicated  interval  for  t  and  value(s)  of  the  constants. 

^  23.  x(?)  =  (acos?,  fcsinr),     0  <  t  <  2jt;    a  =  2,b=l 

^24.  x(r)  =  (2a(l  +  cos  r)  cos  f,  2a(l  +  cos  f)  sin  f),  0< 

t  <  2jt;    a  =  1 

^  25.  x(r)  =  (2a  cosf(l  +  cost)  -  a,  2a  sinf(l  +  cosf)), 
0  <  t  <  2ji;    a  =  l 

^^26.  x(r)  =  (a  sin nt,  b  sin mt),     0  <  t  <  2jt;    a  =  3, 
b  =  2,  n  =  4,  m  =  3 

Find  the  tangential  and  normal  components  of  acceleration  for 
the  paths  given  in  Exercises  27—32. 


(b)  Showthatx(s)=(i(l-.y2),  ^cos"1  s  -  Wl  -  s2 

is  parametrized  by  arclength,  and  compute  its 
curvature. 

In  Exercises  23-26,  (a)  use  a  computer  algebra  system  to  cal- 
culate the  curvature  k  of  the  indicated  path  x  and  (b)  plot  the 


>) 


27.  x(?)  = 

f2i  +  fj 

28.  x(?)  = 

(2t ,  e2') 

29.  x(r)  = 

(e'  cos2f ,  e*  sin  2?) 

30.  x(r)  = 

(4cos5f,  5  sin4f,  30 

31.  x(?)  = 

(t,t,r2) 

32.  x(?)  = 

1(1  —  cos  t  )i  +  sinf  j  +  ^  cos  t  k 

33.  (a)  Show  that  the  tangential  and  normal  compo- 
nents of  acceleration  atang  and  anorm  satisfy  the 
equations 


x  x  x 


^tang 


(b)  Use  these  formulas  to  find  the  tangential  and 
normal  components  of  acceleration  for  the  path 
x(f)  =  (?+2)i  +  f2j  +  3fk 

34.  Use  Exercise  33  to  show  that,  for  the  plane  curve 

y  =  m, 

f'(x)f"(x) 
"tang  —     ,  ) 

yi+(/'«)2 

\f"{*)\ 
^norm  —      ;   ■ 

yi + (f'M)2 

35.  Establish  the  following  formula  for  the  torsion: 

(v  x  a)  •  a' 
II v  x  a[|2 

36.  Show  that  kx  =  — T'  •  B',  where  differentiation  is  with 
respect  to  the  arclength  parameter  s. 

37.  Show  that  if  x  is  a  path  parametrized  by  arclength  and 
x'  x  x"  ^  0,  then 

K2T  =  (X'XX")-X"'. 

38.  Suppose  x:  /  ->  R3  is  a  path  with  x'(f)  x  x"(r)  /  Ofor 
all  t  e  /.  The  osculating  plane  to  the  path  at  t  =  to  is 
the  plane  containing  x(/o)  and  determined  by  (i.e.,  par- 
allel to)  the  tangent  and  normal  vectors  T(fo)  and  N(fo). 


3.3  |  Vector  Fields:  An  Introduction 


The  rectifying  plane  at  t  =  to  is  the  plane  contain- 
ing x(fo)  and  determined  by  the  tangent  and  binormal 
vectors  T(fo)  and  B(?o).  Finally,  the  normal  plane  at 
t  =  to  is  the  plane  containing  x(fo)  and  determined  by 
the  normal  and  binormal  vectors  N(?o)  and  B(fo).  Note 
that  both  the  osculating  and  rectifying  planes  may  be 
considered  to  be  tangent  planes  to  the  path  at  to  since 
they  are  both  parallel  to  T(fo). 

(a)  Show  that  B(fo)  is  perpendicular  to  the  osculating 
plane  at  to,  that  N(?o)  is  perpendicular  to  the  rec- 
tifying plane  at  to,  and  that  T(to)  is  perpendicular 
to  the  normal  plane  at  to- 

(b)  Calculate  the  equations  for  the  osculating,  rec- 
tifying, and  normal  planes  to  the  helix  x(?)  = 
(a  cos  t,  a  sin?,  bi)  at  any  ?o-  (Hint:  To  speed  your 
calculations,  use  the  results  of  Example  9.) 

39.  Recall  that  the  equation  for  a  sphere  of  radius  a  >  0 
and  center  xo  may  be  written  as  ||x  —  xo[|  =  a.  (See 
Example  15  of  §2. 1 .)  Explain  why  the  image  of  a  path 
x  with  the  property  that 

(x(f)  -  x0)  •  (x(r)  -  x0)  =  a2 

for  all  t  must  lie  on  a  sphere  of  radius  a. 

40.  Let  x  be  a  path  with  x'  x  j"  /  0  and  suppose  that  there 
is  a  point  xo  that  lies  on  every  normal  plane  to  x.  Show 
that  the  image  of  x  lies  on  a  sphere.  (See  Exercise  38 
concerning  normal  planes  to  paths.) 

41 .  Use  the  result  of  Exercise  40  to  show  that  x(? )  = 
(cos2f,  —  sin2f,  2 cos/)  lies  on  a  sphere  by  showing 
that  (1,0,0)  lies  on  every  normal  plane  to  x. 

42.  Use  the  result  of  Exercise  27  of  §  1 .4  to  show  that 

N  x  B  =  T       and       B  x  T  =  N. 


As  a  result,  we  can  arrange  T,  N,  and  B  in  a  circle  so 
that  they  correspond,  respectively,  to  the  vectors  i,  j, 
k  appearing  in  Figure  1.54  and  so  that  we  may  use  a 
mnemonic  for  identifying  cross  products  that  is  similar 
to  the  one  described  in  Example  1  of  §  1 .4. 

Let  x  be  a  path  of  class  C3,  parametrized  by  arclength  s, 
with  x'  x  x"  /  0.  We  define  the  Darboux  rotation  vector  (also 
called  the  angular  velocity  vector)  by 

w  =  rT  +  x:B. 

Note  that  w(.?o)  is  parallel  to  the  rectifying  plane  to  x(so).  The 
direction  of  the  Darboux  vector  w  gives  the  axis  of  the  "screw- 
like" motion  of  the  path  x  and  its  length  gives  the  angular 
velocity  of  the  motion.  Exercises  43-45  concern  the  Darboux 
vector. 

43.  Show  that  ||w[|  =  si  k1  +  r2.  (Hint:  The  vectors  T,  N, 
and  B  are  pairwise  orthogonal.) 

44.  (a)  Use  the  Frenet-Serret  formulas  to  establish  the 

Darboux  formulas: 

T'  =  w  x  T 
N'  =  w  x  N 
B'  =  w  x  B. 

(b)  Use  the  Darboux  formulas  to  establish  the  Frenet- 
Serret  formulas.  Hence  the  two  sets  of  equations 
are  equivalent.  (Hint:  Use  Exercise  42.) 

45.  Show  that  x  is  a  helix  if  and  only  if  w  is  a  constant 
vector.  (Hint:  Consider  w'  and  use  Theorem  2.5.) 


/  /  /  / 


Figure  3.28  The  constant  vector 
field  F(x)  =  i  +  j. 


3.3  Vector  Fields:  An  Introduction 

We  begin  with  a  simple  definition. 


DEFINITION  3.1    A  vector  field  on  R"  is  a  mapping 

MC  R"  R". 


We  are  concerned  primarily  with  vector  fields  on  R2  or  R3 .  fn  such  cases,  we 
adopt  the  point  of  view  that  a  vector  field  assigns  to  each  point  x  in  X  a  vector 
F(x)  in  R" ,  represented  by  an  arrow  whose  tail  is  at  the  point  x.  This  perspective 
allows  us  to  visualize  vector  fields  in  a  reasonable  way. 

EXAMPLE  1  Suppose  F:  R2  -■>•  R2  is  defined  by  F(x)  =  a,  where  a  is  a  con- 
stant vector.  Then  F  assigns  a  to  each  point  of  R2,  and  so  we  can  picture  F  by 
drawing  the  same  vector  (parallel  translated  of  course)  emanating  from  each  point 
in  the  plane,  as  suggested  by  Figure  3.28.  ♦ 


Chapter  3  |  Vector-Valued  Functions 


(x,  y) 

G(x,  y) 

(0,0) 

0 

(1,0) 

-j 

(0,1) 

i 

(1,1) 

i- j 

Figure  3.29  The  vector  field 
G(x,  y)  =  yi  —  xj  of  Example  2. 


EXAMPLE  2  Let's  depict  G:  R2  ->  R2,  G(x,  y)  =  yi-  xj.  We  can  begin  to  do 
this  by  calculating  some  specific  values  of  G,  as  in  the  adjacent  table.  However, 
it  is  difficult  to  get  much  of  a  feeling  for  G  as  a  whole  in  this  way.  To  understand 
G  somewhat  better,  we  need  to  "play  around"  a  bit.  Note  that 


IGCoOII  =  llvi-xjl 


=  r 


where  r  =  xi  +  yj,  the  position  vector  of  the  point  (x,  y).  From  this  observation, 
it  follows  that  G  has  constant  length  a  on  the  circle  x2  +  y2  =  a2.  In  addition, 
we  have 


r  •  G(x,  y)  =  (xi  +  yj)  •  (yi  -  xj)  =  0. 

Hence,  G(x,  y)  is  always  perpendicular  to  the  position  vector  of  the  point  (x,  y). 
These  facts,  together  with  a  table  like  the  preceding  one,  make  it  possible  to  see 
that  G  looks  like  Figure  3.29.  ♦ 


Remark  Sometimes  a  scalar- valued  function  f:X  c  R"  — >•  R  is  called  a  scalar 
field.  One  thinks  of  a  vector  field  on  R"  as  attaching  vector  information  (such 
as  wind  velocity)  to  each  point  and  a  scalar  field  as  attaching  real  number  infor- 
mation (such  as  temperature  or  pressure).  We'll  use  the  term  "scalar  field"  only 
occasionally,  but  we  don't  want  to  shock  you  when  we  do. 

EXAMPLE  3  Let  r  =  xi  +  yj  +  zk.  The  so-called  inverse  square  vector  field 
in  R3  is  a  function  F:  R3  —  {0}  — >  R3  given  by 


F(x,  y,  z)  = 


where  c  is  any  (nonzero)  constant.  If  the  term  "inverse  square"  seems  inappropriate 
to  you,  we'll  try  to  convince  you  otherwise.  Setu  =  r/||r||  sothatr  =  ||r||u.  Then 
F  is  given  by 


F(x,  y,  z)  = 


Figure  3.30  An  inverse  square 
vector  field. 


(1) 


Therefore,  F  is  a  vector  field  whose  direction  at  the  point  P(x,  y,  z)  ^  (0,  0,  0) 
is  parallel  to  the  vector  from  the  origin  to  P  and  whose  magnitude  is  inversely 
proportional  to  the  square  of  the  distance  from  the  origin  to  P.  Note  that  F  points 
away  from  the  origin  if  c  is  positive  and  toward  the  origin  if  c  is  negative. 

We  have  seen  an  example  of  an  inverse  square  field  in  §3.1 — namely,  the 
Newtonian  gravitational  field  between  two  bodies.  If  one  of  the  bodies  is  at  the 
origin  and  the  other  at  (x,  y,  z),  then  we  have 

GMm 
F=  ^u. 


In  this  case,  the  proportionality  constant  c  is  —GMm,  which  is  negative.  This 
means  that  the  gravitational  force  is  attractive  (i.e.,  it  points  in  the  direction 
that  reduces  the  distance  between  the  two  bodies).  Such  a  vector  field  is  shown  in 
Figure  3 .3  0.  An  example  of  a  repelling  inverse  square  field  is  the  electrostatic  force 


3.3  |  Vector  Fields:  An  Introduction 


between  two  particles  with  like  static  charges  (both  positive  or  both  negative). 
This  force  is  expressed  by  Coulomb's  law, 

F  =  z-  u, 


where  r  is  the  vector  from  particle  1  (at  the  origin)  to  particle  2,  u  =  r/||r||, 
q\  and  qi  are  the  respective  charges  (positive  or  negative)  on  the  particles,  and 
k  is  a  constant  appropriate  for  the  units  being  used.  In  mks  units,  distance  is 
measured  in  meters,  charge  in  coulombs,  force  in  newtons,  so  that  k  is  equal  to 
8.9875  x  109  Nm2/C2.  ♦ 


Gradient  Fields  and  Potentials 

Inverse  square  fields  are  interesting  not  only  for  their  origin  in  basic  physical 
situations,  but  also  because  they  are  examples  of  gradient  fields.  A  gradient  field 
on  R"  is  a  vector  field  F:  X  c  R"  — >  R"  such  that  F  is  the  gradient  of  some 
(differentiable)  scalar- valued  function  /:  X  ->  R.  That  is, 

F(x)  =  V/(x) 

at  all  x  in  X.  The  function  /  is  called  a  (scalar)  potential  function  for  the  vector 
field  F.  To  see  what  this  means  in  the  case  of  the  inverse  square  field  (1),  we  write 
out  the  components  of  F  explicitly: 

xi  +  yj  +  zk  \ 


||r||2         V*2  +  y2  +  z2J  \jx2  +  y2  +  z2 
since  r  =  xi  +  yj  +  zk  and  u  =  r/||r||.  That  is, 

(X'  y'  Z)  ~  (x2  +  y2  +  z2)3/2  1  +  (x2  +  y2  +  z2)3/2  J  +  (x2  +  y2  +  z2)3/2  ' 

We  leave  it  to  you  to  check  that  F(x,  y,  z)  =  V/(x,  y,  z),  where /:R3  —  {0}  ->  R 
is  given  by 


f(x,  y,z) 


c 


y/x2  +  y2  +  z2  Ilrll 


Remark  In  physics  and  engineering,  a  negative  sign  is  often  introduced  in  the 
definition  of  a  potential  function  (i.e.,  so  that  a  potential  function  g  for  a  vector 
field  F  is  one  such  that  F  =  —  Vg).  The  motivation  behind  such  a  convention  is 
that  in  physical  applications,  it  is  desirable  to  have  the  potential  function  rep- 
resent potential  energy  in  some  sense.  For  example,  in  the  case  of  the  gravita- 
tional field  F  =  —  ( G  M/w  / 1 1  r  1 1 2  )u,  a  physicist  would  take  the  potential  function  to 
be  —GMm/\\r\\,  not  +GMm/||r||  as  we  do.  The  advantage  to  the  physicist  in 
doing  so  is  that  the  physicist's  potential  function  increases  with  increasing  ||r||. 
This  corresponds  to  the  notion  that  the  greater  the  distance  between  two  bodies, 
the  greater  should  be  the  stored  gravitational  potential  energy. 

From  Theorem  6.4  of  Chapter  2  we  know  that  the  gradient  of  any  C1  scalar- 
valued  function  /:  X  c  R"  ->•  R  is  perpendicular  to  the  level  sets  of  /.  Thus,  if 
F  is  a  gradient  vector  field  on  R" ,  F(x)  must  be  perpendicular  to  the  level  set  of  a 


224       Chapter  3  I  Vector- Valued  Functions 


Figure  3.31  A  gradient  vector  field  F  =  V/.  Equipotential 
lines  are  shown  where  /  is  constant. 


potential  function  of  F  containing  the  point  x.  If  /  is  such  a  potential  function,  the 
level  set  {x  |  f(x)  =  c]  is  called  an  equipotential  set  (or  equipotential  surface  if 
n  =  3,  or  equipotential  line  if  n  =  2)  of  the  vector  field  F.  (See  Figure  3.31.) 

You've  seen  examples  of  equipotential  lines  every  time  you've  looked  at  a 
weather  map.  Usually  curves  of  constant  barometric  pressure  (called  isobars)  or 
of  constant  temperature  (isotherms)  are  drawn.  (  See  Figure  3.32.)  Perpendicular 
to  such  equipotential  lines  are  associated  gradient  vector  fields  that  point  in  the 
direction  of  most  rapid  increase  of  pressure  or  temperature. 

Flow  Lines  of  Vector  Fields 

When  you  draw  a  sketch  of  a  vector  field  on  R2  or  R3,  it  is  easy  to  imagine 
that  the  arrows  represent  the  velocity  of  some  fluid  moving  through  space  as  in 
Figure  3.33.  It's  natural  to  let  the  arrows  blend  into  complete  curves.  What  you're 


Figure  3.32  A  weather  map.  (Weather  graphics  courtesy  of  Accuweather,  Inc.  385 
Science  Park  Road,  State  College,  PA  16803.  (814)  237-0309.  ©  2011.  Used  with 
permission.) 


Figure  3.33  A  fluid  moving 
through  space. 


3.3  I  Vector  Fields:  An  Introduction 


225 


doing  analytically  is  drawing  paths  whose  velocity  vectors  coincide  with  those  of 
the  vector  field. 


Figure  3.34  A  flow  line. 


Figure  3.35  The  vector  field 
F(x,>>,z)  =  2i-3j  +  kof 
Example  4. 


Figure  3.36  Flow  lines  of 

F(x,  y)  =  —yi  +  x\  of  Example  5. 


DEFINITION  3.2  A  flow  line  of  a  vector  field  F:  X  C  R"  ->  R"  is  a  differ- 
entiable  path  x:  /  — >  R"  such  that 

x'(t)  =  F(x(0). 

That  is,  the  velocity  vector  of  x  at  time  ?  is  given  by  the  value  of  the  vector 
field  F  at  the  point  on  x  at  time  ?.  (See  Figure  3.34.) 


EXAMPLE  4  We  calculate  the  flow  lines  of  the  constant  vector  field 
F(x,y,z)  =  2i-3j  +  k. 

A  picture  of  this  vector  field  (see  Figure  3.35)  makes  it  easy  to  believe  that 
the  flow  lines  are  straight-line  paths.  Indeed  if  x(?)  =  (x(t),  y(t),  z(t))  is  a  flow 
line,  then,  by  Definition  3.2,  we  must  have 

x'(?)  =  (x'(?),  y'(t),  z'{t))  =  (2,  -3,  1)  =  F(x(?)). 

Equating  components,  we  see 

V(0  =  2 
y'(t)  =  -3. 
z'(t)=  1 

These  differential  equations  are  readily  solved  by  direct  integration;  we  obtain 

x(?)  =  2?  +  xq 
y{t)  =  -3?  +  y0  , 
z(t)  =  t  +  zo 

where  xq,  yo,  and  zo  are  arbitrary  constants.  Hence,  as  expected  we  obtain  para- 
metric equations  for  a  straight-line  path  through  an  arbitrary  point  (x0,  yo,  Zo) 
with  velocity  vector  (2,-3,1).  ♦ 

EXAMPLE  5  Your  intuition  should  lead  you  to  suspect  that  a  flow  line  of  the 
vector  field  F(x,  y)  =  — yi  +  xj  should  be  circular  as  showninFigure  3.36.  Indeed 
if  x:  [0,  2jt)  — >  R2  is  given  by  x(t)  =  (a  cos  t ,  a  sin  t),  where  a  is  constant,  then 

x'(f)  =  —  a  sinf  i  +  a  cost  j  =  F(a  cos/,  a  sin?), 

so  such  paths  are  indeed  flow  lines. 

Finding  all  possible  flow  lines  of  F(x ,  y)  =  —  yi  +  xj  is  a  more  involved  task. 
If  x(/)  =  (x(f),  y(t))  is  a  flow  line,  then,  by  Definition  3.2,  we  must  have 

At)  =  x'(t)i  +  y'(t)j  =  -y(0i  +  *(0j  =  F(x(0). 
Equating  components, 

\x'{t)  =  -y(t) 
y'(t)  =  x(t)  ■ 

This  is  an  example  of  a  first-order  system  of  differential  equations.  It  turns  out 
that  all  solutions  to  this  system  are  of  the  form 


x(t)  =  (a  cos  t  —  b  sin?,  a  sin?  +  b  cos  ?), 


Chapter  3  |  Vector-Valued  Functions 


where  a  and  b  are  arbitrary  constants.  It's  not  difficult  to  see  that  such  paths  trace 
circles  when  at  least  one  of  a  or  b  is  nonzero.  ♦ 

In  general,  if  F  is  a  vector  field  on  R",  finding  the  flow  lines  of  F  is  equivalent 
to  solving  the  first-order  system  of  differential  equations 

x[(t)  =  FiCti(f),  X2(t),  X„(t)) 
x'2(t)  =  F2(xi(t),  x2(t),  x„(t)) 

.  X'„(t)  =  Fn(Xi(t),  X2(t),  .  .  .  ,  Xn(t)) 

for  the  functions  X\(t),  . . . ,  xn(t)  that  are  the  components  of  the  flow  line  x.  (The 
function  F,  is  just  the  ith  component  function  of  the  vector  field  F.)  Such  a 
problem  takes  us  squarely  into  the  realm  of  the  theory  of  differential  equations,  a 
fascinating  subject,  but  not  of  primary  concern  at  the  moment. 


3.3  Exercises 


In  Exercises  1-6,  sketch  the  given  vector  fields  on  R2. 


1. 

F 

=  yi-X\ 

2. 

F 

=  xi-  yj 

3. 

F 

=  {-*,y) 

4. 

F 

=  (x,x2) 

5. 

F 

=  (x2,x) 

6. 

F 

=  (y\y) 

In  Exercises  7-12,  sketch  the  given  vector  field  on  R3. 

7.  F  =  3i  +  2j  +  k 

8.  F  =  (y,  -x,0) 

9.  F  =  (0,  z,  -y) 

10.  F  =  (y,  -x,2) 

11.  F  =  (y,-x,z) 

V  x 

12.  F=    -      y  =  i  — =  j 

V-*2  +  y2  +  z2      V-*2  +  y2  +  z2 

7 

+     ;         =  =k 

s/x2  +  y2  +  z2 

In  Exercises  13-16,  use  a  computer  to  plot  the  given  vector 
fields  over  the  indicated  ranges. 

^  13.  F  =  (x  -  y,  x  +  y);     -1  <  x  <  1,  -1  <  y  <  1 
^14.  F  =  (y3x,  x2y);     -2  <  x  <  2,  -2  <  y  <  2 

15.  F  =  (x  siny,  y  cosx);  —2jt<x<2jt, 
—2it  <  y  <  2tc 

^  16.  F  =  (cos(x  -  y),  sin(x  +  y));     -2n  <  x  <  2jt, 
—2it  <  y  <  2tc 


In  Exercises  17-19,  verify  that  the  path  given  is  a  flow  line  of 
the  indicated  vector  field.  Justify  the  result  geometrically  with 
an  appropriate  sketch. 

17.  x(t)  =  (sinf,  cosr,  0),  F  =  (y,  -x,  0) 

18.  x(r)  =  (sinf,  cos?,  2t),  F  =  (y,  -x,  2) 

19.  x(r)  =  (sinf,  cosf,  e2t),  F  =  (y,  -x,  2z) 

In  Exercises  20-22,  calculate  the  flow  line  x(f )  of  the  given 
vector  field  F  that  passes  through  the  indicated  point  at  the 
specified  value  of  t. 

20.  F(x,y)  =  -xi  +  yj;    x(0)  =  (2,  1) 

21.  F(x,y)  =  (x2,y);     x(l)  =  (l,e) 

22.  F(x,y,z)  =  2i-3yj  +  Z3k;    x(0)  =  (3,5,7) 

23.  Consider  the  vector  field  F  =  3  i  -  2  j  +  k. 

(a)  Show  that  F  is  a  gradient  field. 

(b)  Describe  the  equipotential  surfaces  of  F  in  words 
and  with  sketches. 

24.  Consider  the  vector  field  F  =  2x  i  +  2y  j  —  3k. 

(a)  Show  that  F  is  a  gradient  field. 

(b)  Describe  the  equipotential  surfaces  of  F  in  words 
and  with  sketches. 

25.  If  x  is  a  flow  line  of  a  gradient  vector  field  F  =  V/, 
show  that  the  function  G(f )  =  /(x(f ))  is  an  increasing 
function  of  f.  (Hint:  Show  that  G'(t)  is  always  non- 
negative.)  Thus,  we  see  that  a  particle  traveling  along 
a  flow  line  of  the  gradient  field  F  =  V/  will  move 
from  lower  to  higher  values  of  the  potential  function 
/.  That's  why  physicists  define  a  potential  function  of 
a  gradient  vector  field  F  to  be  a  function  g  such  that 
F  =  —  Vg  (i.e.,  so  that  particles  traveling  along  flow 
lines  move  from  higher  to  lower  values  of  g). 


3.4  |  Gradient,  Divergence,  Curl,  and  the  Del  Operator  227 


LetF:  X  C  R"  — >  R"  be  a  continuous  vector  field.  Let  (a,  b)be 
an  interval  in  "Rthat  contains  0.  (Think  of  {a,  b)asa  "timeinter- 
val")  A  flow  ofF  is  a  differentiable  function  cp:  X  x  (a,  b)  — > 
R"  ofn  +  1  variables  such  that 


dt 


0(x,  t)  =  F(0(x,  f));    0(x,  0)  =  x. 


Intuitively,  we  think  of  0(x,  t)  as  the  point  at  time  t  on  the  flow 
line  ofF  that  passes  through  x  at  time  0.  (See  Figure  3.37.) 
Thus,  the  flow  ofF  is,  in  a  sense,  the  collection  of  all  flow  lines 
ofF.  Exercises  26-31  concern  flows  of  vector  fields. 


F(0(x,  0) 
<Hx,f)/ 


Figure  3.37  The  flow  of  the  vector  field  F. 


26.  Verify  that 

0:R2  x  R^  R2, 

<j>(x,  .v,  t) 


x  +  y  t  ,x~y  -t 

 e  -\  e  , 

2  2 


x  +  y   ,     y  —  x 

  e'  +  :  e~ 

2  2 

is  a  flow  of  the  vector  field  F(x,  y)  =  (y,  x). 


27.  Verify  that 
</>:R2  x  R^  R2, 

4>(x,  y,  t)  =  (y  sinf  +  x  cos  t,  y  cos?  —  x  sinf) 
is  a  flow  of  the  vector  field  F(x,  y)  =  (y,  —x). 

28.  Verify  that 
0:R3  x  R^  R3, 

4>(x,  y,  z,  t)  =  (x  cos2f  —  y  sin 2;,  y  cos2f 
+  x  sin2f,  ze~') 


is  a  flow  of  the  vector  field  F(x,y,z) 
2x  j  —  z  k. 


-2vi  + 


29.  Show  that  if  <j>:  X  x  (a,  fo)  R"  is  a  flow  ofF,  then, 
for  a  fixed  point  xo  in  X,  the  map  x:  (a,  b)  — ►  R"  given 
by  x(f)  =  0(xo,  f)  is  a  flow  line  ofF. 

30.  If  0  is  a  flow  of  the  vector  field  F,  explain  why 
0(0(x,  r),  s)  =  0(x,  s  +  r).  (Hint:  Relate  the  value  of 
the  flow  4>  at  (x,  t)  to  the  flow  line  ofF  through  x.  You 
may  assume  the  fact  that  the  flow  line  of  a  continuous 
vector  field  at  a  given  point  and  time  is  determined 
uniquely.) 

31 .  Derive  the  equation  of  first  variation  for  a  flow  of  a 
vector  field.  That  is,  if  F  is  a  vector  field  of  class  C1 
with  flow  <p  of  class  C2,  show  that 


dt 


Dx0(x,  0  =  DF(0(x,  t))D%<p(x,  t). 


Here  the  expression  "Dx0(x,  t)"  means  to  differentiate 
(p  with  respect  to  the  variables  x\ ,  X2, . . . ,  x„,  that  is, 
by  holding  t  fixed. 


3.4  Gradient,  Divergence,  Curl,  and  the  Del 
Operator 

In  this  section,  we  consider  certain  types  of  differentiation  operations  on  vector 
and  scalar  fields.  These  operations  are  as  follows: 

1.  The  gradient,  which  turns  a  scalar  field  into  a  vector  field. 

2.  The  divergence,  which  turns  a  vector  field  into  a  scalar  field. 

3.  The  curl,  which  turns  a  vector  field  into  another  vector  field.  (Note:  The  curl 
will  be  denned  only  for  vector  fields  on  R3.) 

We  begin  by  defining  these  operations  from  a  purely  computational  point  of  view. 
Gradually,  we  shall  come  to  understand  their  geometric  significance. 


The  Del  Operator  

The  del  operator,  denoted  V,  is  an  odd  creature.  It  leads  a  double  life  as  both 
differential  operator  and  vector.  In  Cartesian  coordinates  on  R3,  del  is  defined  by 


Chapter  3  |  Vector-Valued  Functions 

the  curious  expression 


The  "empty"  partial  derivatives  are  the  components  of  a  vector  that  awaits  suitable 
scalar  and  vector  fields  on  which  to  act.  Del  operates  on  (i.e.,  transforms)  fields 
via  "multiplication"  of  vectors,  interpreted  by  using  partial  differentiation. 

For  example,  if  /:  X  c  R3  — >•  R  is  a  differentiable  function  (scalar  field), 
the  gradient  of  /  may  be  considered  to  be  the  result  of  multiplying  the  vector  V 
by  the  scalar  /,  except  that  when  we  "multiply"  each  component  of  V  by  /,  we 
actually  compute  the  appropriate  partial  derivative: 

/  3        3         3  \  df      df  df 

Vf(x,y,z)=    i—  +j— +k-  )f(x,y,z)=  -M+/j  +  /k. 

\  dx       ay        dz )  dx       dy  dz 

The  del  operator  can  also  be  defined  in  R",  for  arbitrary  n.  If  we  take 
x\,  X2,  ■  ■  ■ ,  xn  to  be  coordinates  for  R",  then  del  is  simply 


where  e, ■  =  (0, . . . ,  1, . . . ,  0),  i  =  1, . . . ,  n,  is  the  standard  basis  vector  for  R". 
The  Divergence  of  a  Vector  Field  


Whereas  taking  the  gradient  of  a  scalar  field  yields  a  vector  field  the  process  of 
taking  the  divergence  does  just  the  opposite:  It  turns  a  vector  field  into  a  scalar 
field. 


DEFINITION  4.1  Let  F:  X  C  R"  R"  be  a  differentiable  vector  field. 
Then  the  divergence  of  F,  denoted  div  F  or  V  •  F  (the  latter  read  "del  dot 
F"),  is  the  scalar  field 

dFi      dF2  dFn 
div  F  =  V  •  F  =  — L  +  — L +  ...  +  — 1, 

3xi      dxi  dx„ 

where  Cartesian  coordinates  for  R"  and  F\ , . . . ,  F„  are  the 

component  functions  of  F. 


It  is  essential  that  Cartesian  coordinates  be  used  in  the  formula  of  Definition  4.1. 
(Later  in  this  section  we  shall  see  what  div  F  looks  like  in  cylindrical  and  spherical 
coordinates  for  R3 .) 

EXAMPLE  1     IfF  =  jr2vi  +  xZj+xyzk,then 

3  3d 
div  F  =  —  (x2y)  +  — (jcz)  +  —  (xyz)  =  2xy  +  0  +  xy  =  3xy.  ♦ 
dx  dy  dz 


3.4  |  Gradient,  Divergence,  Curl,  and  the  Del  Operator  229 


The  notation  for  the  divergence  involving  the  dot  product  and  the  del  operator 
is  especially  apt:  If  we  write 


F  =  Fid  +  F2e2  H  h  Fnen, 


then, 


/     3  3  3  \  , 

V-F=    ei—  +e2—  +  ---  +  e„—    ■  (Fid  +  F2e2  +  ■  ■  ■  +  F„e„) 

\     dXl  0X2  oxnJ 

_  dFi      dF2  dFn 
dx\      9x2  3x„ ' 

where,  once  again,  we  interpret  "multiplying"  a  function  by  a  partial  differential 
operator  as  performing  that  partial  differentiation  on  the  given  function. 

Intuitively,  the  value  of  the  divergence  of  a  vector  field  at  a  particular  point 
gives  a  measure  of  the  "net  mass  flow"  or  "flux  density"  of  the  vector  field  in 
or  out  of  that  point.  To  understand  what  such  a  statement  means,  imagine  that 
the  vector  field  F  represents  velocity  of  a  fluid.  If  V  •  F  is  zero  at  a  point,  then 
the  rate  at  which  fluid  is  flowing  into  that  point  is  equal  to  the  rate  at  which 
fluid  is  flowing  out.  Positive  divergence  at  a  point  signifies  more  fluid  flowing  out 
than  in,  while  negative  divergence  signifies  just  the  opposite.  We  will  make  these 
assertions  more  precise,  even  prove  them,  when  we  have  some  integral  vector 
calculus  at  our  disposal.  For  now,  however,  we  remark  that  a  vector  field  F  such 
that  V  •  F  =  0  everywhere  is  called  incompressible  or  solenoidal. 

EXAMPLE  2   The  vector  field  F  =  xi  +  yj  has 

V-F=  A(je)+  ^(y)  =  2. 
dx  ay 

This  vector  field  is  shown  in  Figure  3.38.  At  any  point  in  R2,  the  arrow  whose 
tail  is  at  that  point  is  longer  than  the  arrow  whose  head  is  there.  Hence,  there  is 
greater  flow  away  from  each  point  than  into  it;  that  is,  F  is  "diverging"  at  every 
point.  (Thus,  we  see  the  origin  of  the  term  "divergence.") 

The  vector  field  G  =  — xi  —  yj  points  in  the  direction  opposite  to  the  vector 
field  F  of  Figure  3.38  (see  Figure  3.39),  and  it  should  be  clear  how  G's  divergence 
of  —2  is  reflected  in  the  diagram.  ♦ 


EXAMPLE  3  The  constant  vector  field  F(x,  y,  z)  =  a  shown  in  Figure  3.40 
is  incompressible.  Intuitively,  we  can  see  that  each  point  of  R3  has  an  arrow 
representing  a  with  its  tail  at  that  point  and  another  arrow,  also  representing  a, 
with  its  head  there. 

The  vector  field  G  =  yi  —  xj  has 

v-g=  A(v)  + A(_x)  =  o. 

ax  ay 

A  sketch  of  G  reveals  that  it  looks  like  the  velocity  field  of  a  rotating  fluid,  without 
either  a  source  or  a  sink.  (See  Figure  3.41 .)  ♦ 


The  Curl  of  a  Vector  Field 

If  the  gradient  is  the  result  of  performing  "scalar  multiplication"  with  the  del 
operator  and  a  scalar  field,  and  the  divergence  is  the  result  of  performing  the 
"dot  product"  of  del  with  a  vector  field,  then  there  seems  to  be  only  one  simple 


Chapter  3  |  Vector-Valued  Functions 


Figure  3.40  The  constant  vector 
field  F  =  a. 


Figure  3.41  The  vector  field 
G  =  yi  —  x']  resembles  the 
velocity  field  of  a  rotating  fluid. 


differential  operation  left  to  be  built  from  del.  We  call  it  the  curl  of  a  vector  field 
and  define  it  as  follows: 


DEFINITION  4.2  Let  F:  X  c  R3  ->  R3  be  a  differentiable  vector  field  on 
R3  only.  The  curl  of  F,  denoted  curl  F  or  V  x  F  (the  latter  read  "del  cross 
F"),  is  the  vector  field 

(3  3  3  \ 

V  +  V  +  V )  x     +  F^  +  F3k> 
dx      ay       dz ) 


i  j  k 

d/dx      d/dy  d/dz 
F\         F2  F3 


^EL  -  i  +  ( ?H  _  ^fl\  ■  +  ( ^fl  _  | 

dy       dz  J       \  dz       dx  J       \  dx  dy 


There  is  no  good  reason  to  remember  the  formula  for  the  components  of  the 
curl — instead  simply  compute  the  cross  product  explicitly. 

EXAMPLE  4   If  F  =  x2yi  -  2xzj  +  (x  +  y  -  z)k,  then 


V  x  F 


i  j  k 

d/dx       d/dy  d/dz 

—  2xz      x  +  y  —  z 


x2y 


3  3  \       (  3  3 

—  (x  +  y-z)-  ^~2xz))  1  +  \dz~(x2y)  ~  9^(X  +  y  ~  Z- 


3 


9 


=  (1  +  2x)i  -  j  -  (x2  +  2z)k. 


3.4  |  Gradient,  Divergence,  Curl,  and  the  Del  Operator  231 


Figure  3.42  A  twig  in  a  pond  where  water  moves  with  velocity  given  by  a  vector  field  F.  In  the  left  figure,  the  twig  does 
not  rotate  as  it  travels,  so  curl  F  =  0.  In  the  right  figure,  curl  F  /  0,  since  the  twig  rotates. 


One  would  think  that,  with  a  name  like  "curl,"  V  x  F  should  measure  how 
much  a  vector  field  curls.  Indeed  the  curl  does  measure,  in  a  sense,  the  twisting 
or  circulation  of  a  vector  field  but  in  a  subtle  way:  Imagine  that  F  represents  the 
velocity  of  a  stream  or  lake.  Drop  a  small  twig  in  the  lake  and  watch  it  travel. 
The  twig  may  perhaps  be  pushed  by  the  current  so  that  it  travels  in  a  large  circle, 
but  the  curl  will  not  detect  this.  What  curl  F  measures  is  how  quickly  and  in  what 
orientation  the  twig  itself  rotates  as  it  moves.  (See  Figure  3.42.)  We  prove  this 
assertion  much  later,  when  we  know  something  about  line  and  surface  integrals. 
For  now,  we  simply  point  out  some  terminology:  A  vector  field  F  is  said  to  be 
irrotational  if  V  x  F  =  0  everywhere. 


EXAMPLE  5    Let  F  =  (3x2z  +  y2)  i  +  2xy  j  +  (x3  -  2z)  k.  Then 


V  x  F 


l 

d/dx 
3x2z  +  y2 

a 


j 

d/dy 
2xy 


k 

d/dz 


2z 


dy 


(x>  -  2z) 


^(2xy)W(^(3x2z  +  y2) 


dx 


(x3-2z) 


dx 


(2xy) 


3 

9y 


(3*2  z  +  y2)  I  k 


=  (0  -  0)i  +  (3x2  -  3x2)\  +  (2y  -  2y)k  =  0. 


Thus,  F  is  irrotational. 


Two  Vector-analytic  Results   

It  turns  out  that  the  vector  field  F  in  Example  5  is  also  a  gradient  field.  Indeed 
F  =  V/,  where  f(x,  y,  z)  =  x3z  +  xy2  —  z2.  (We'll  leave  it  to  you  to  verify 
this.)  In  fact,  this  is  not  mere  coincidence  but  an  illustration  of  a  basic  result 
about  scalar-valued  functions  and  the  del  operator: 


THEOREM  4.3  Let  /:  X  C  R3  ->  R  be  of  class  C2.  Then  curl  (grad/)  =  0. 
That  is,  gradient  fields  are  irrotational. 


Chapter  3  |  Vector-Valued  Functions 


PROOF  Using  the  del  operator,  we  rewrite  the  conclusion  as 


V  x  (V/)  =  0, 

which  might  lead  you  to  think  that  the  proof  involves  nothing  more  than  noting 
that  V/  is  a  "scalar"  times  V,  hence,  "parallel"  to  V,  so  thatthe  cross  productmust 
be  the  zero  vector.  However,  V  is  not  an  ordinary  vector,  and  the  multiplications 
involved  are  not  the  usual  ones.  A  real  proof  is  needed. 

Such  a  proof  is  not  hard  to  produce:  We  need  only  start  calculating  V  x  (V/). 
We  have 


df  df  df 
dx       dy  dz 


Therefore, 


V  x  (V/)  = 


i  j  k 

d/dx  d/dy  d/dz 
df/dx      df/dy  df/dz 


d2f 

dydz  dzdy 


dzdy)  \ 


d2f  d2f 


dzdx  dxdz 


j  + 


d2f  d2f 


dxdy  dydx 


Since  /  is  of  class  C2,  we  know  that  the  mixed  second  partials  don't  depend 
on  the  order  of  differentiation.  Hence,  each  component  of  V  x  (V/)  is  zero,  as 
desired.  ■ 

There  is  another  result  concerning  vector  fields  and  the  del  operator  that  is 
similar  to  Theorem  4.3: 


THEOREM  4.4  Let  F:XcR3^R3  be  a  vector  field  of  class  C2.  Then 
div  (curl  F)  =  0.  That  is,  curl  F  is  an  incompressible  vector  field. 


The  proof  is  left  to  you. 

EXAMPLE  6   If  F  =  (xz  —  e2x  cos  z)  i  -  yz  j  +  e2l(sin  y  +  2  sin  z)  k,  then 

3  3  3 

V  •  F  =  —  (xz  —  e2x  cosz)  +  —  (-yz)  +  —  (e2x(smy  +  2sinz)) 
dx  dy  dz 

=  z  -  2e2x  cos  z  -  z  +  2e2x  cos  z  =  0 

for  all  (ij,z)£R5.  Hence,  F  is  incompressible.  We'll  leave  it  to  you  to  check 
that  F  =  V  x  G,  where  G(x,  y,  z)  =  e2x  cos  y  i  +  e2x  sin  z  j  +  xyz  k,  so  that,  in 
view  of  Theorem  4.4  the  incompressibility  of  F  is  not  really  a  surprise.  ♦ 


Other  Coordinate  Formulations  (optional) 

We  have  introduced  the  gradient,  divergence,  and  curl  by  formulas  in  Cartesian 
coordinates  and  have,  at  least  briefly,  discussed  their  geometric  significance.  Since 
certain  situations  may  necessitate  the  use  of  cylindrical  or  spherical  coordinates, 
we  next  list  the  formulas  for  the  gradient,  divergence,  and  curl  in  these  coordinate 
systems.  Before  we  do,  however,  a  remark  about  notation  is  in  order.  Recall  that 
in  cylindrical  coordinates,  there  are  three  unit  vectors  er,  e$,  and  ez  that  point  in 
the  directions  of  increasing  r,  6,  and  z  coordinates,  respectively.  Thus,  a  vector 


3.4  |  Gradient,  Divergence,  Curl,  and  the  Del  Operator  233 


field  F  on  R3  may  be  written  as 

F  =  Frer  +  Fgee  +  F-e.. 

In  general,  the  component  functions  Fr,  Fg,  and  Fz  are  each  functions  of  the 
three  coordinates  r,  9,  and  z;  the  subscripts  serve  only  to  indicate  to  which  of 
the  vectors  e, ,  eg,  and  e;  that  particular  component  function  should  be  attached. 
Similar  comments  apply  to  spherical  coordinates,  of  course:  There  are  three  unit 
vectors  ep,  e9,  and  eg,  and  any  vector  field  F  can  be  written  as 

F  =  Fpep  +  Fvev  +  Feeg. 


THEOREM  4.5    Let  /:  X  c  R3  ->  R  and  F:  Y  C  R3 

scalar  and  vector  fields,  respectively.  Then 


3/        13/     ,  3/ 

— er  H  ee  H  e,; 

9r        r  30  3z 


divF  = 


curl  F  = 


3  3F0  3 

or  39  dz 


R  be  differentiate 

(3) 
(4) 


er  rt0 

3/3r  3/36* 
Fr  rFg 


e 

3/3z 


(5) 


PROOF  We'll  prove  formula  (4)  only,  since  the  argument  should  be  sufficiently 
clear  so  that  it  can  be  modified  to  give  proofs  of  formulas  (3)  and  (5).  The  idea  is 
simply  to  rewrite  all  rectangular  symbols  in  terms  of  cylindrical  ones. 
From  the  equations  in  (8)  of  §  1 .7,  we  have 


e,-  =  cos  6  i  +  sin  9  j 
eg  =  —  sin  9  i  +  cos  6  j . 
e-  =  k 


(6) 


From  the  chain  rule,  we  have  the  following  relations  between  rectangular  and 
cylindrical  differential  operators: 

3  3  3 

—  =  cos  9  h  sin  6  — 

3r  dx  3y 

3  3  3 

—  =  —rsm9  h  r  cos  9 —  . 

89  dx  3y 

3  _  3 

dz  dz 

These  relations  can  be  solved  algebraically  for  d/dx,  d/dy,  and  3/3z  to  yield 


a 

=  cos 

3 

9  

sin  9 

3 

37 

dr 

r 

39 

3 

=  sin  6 

3 

Vr  + 

cos  9 

8 

3v" 

r 

39 

3 

3 

dz 

~  d~Z 

(7) 


234       Chapter  3  I  Vector- Valued  Functions 


Hence,  we  can  use  (6)  and  (7)  to  rewrite  the  expression  for  the  divergence  of  a 
vector  field  on  R3 : 


V-F 


3         3  3 

IT  1  +  IT  J  +  IT  k  )  *  (F'"e'-  +  FsCs  +  F^ 
dx       dy  dz 


1  cos 


,1 

dr 


sin#  3 
~rr~d~9 


j    sin  (9 


3      cos9  3 


8r 


r  3(9 


dz 


[(Fr  cos  9  -  Fe  sin  9)  i  +  (Fr  sin  <9  +  Fe  cos  0)  j  +  Fz  k]. 


(We  used  the  equations  in  (7)  to  rewrite  the  partial  operators  d/dx,  3/3 y,  and 
3/3z  appearing  in  del  and  the  equations  in  (6)  to  replace  the  cylindrical  basis 
vectors  er,  eg,  and  ez  by  expressions  involving  i,  j,  and  k.)  Performing  the  dot 
product  and  using  the  product  rule  yields 


V-F 


cost 


3      sva9  3 

i  

dr        r  d9 

3      cos#  3 

+  |sin6>  1  

dr        r  d9 


(Fr  cos  9  —  Fg  sin#) 


(Fr  sin  9  +  Fe  cos  0)  H  F- 

dz 


dFr  3 

=  cost/  [cos 9  h  Fr  —  (cos#) 

dr  dr 


cos  v  sin 


3r 

dFr 


dr 


sin9 
r 

sin 


+  sm9  \  sm9 — -  +  Fr  —  (sm.9) 
dr  dr 

dF0  3 

+  sin#  | cosf  h  Fg — (cos$) 

dr  dr 


dFr  3 

cos  9  h  Fr  —  (cos  9) 

3(9  3(9 


r 

cos  9 
r 

cos  9 


sin  9- 


dFe 
d9 


3 

Fe—(sin9) 
3(9 v 


dFr  d 
sm9 — -  +  Fr— (sine) 
36>  89 


cos9- 


d_F^ 
89 


3 

Fa — (cos  9) 
d9 


+ 


dz 


After  some  additional  algebra,  we  find  that 


3  F 

V  •  F  =  (cos2  9  +  sin2  9) — -  + 

dr 


sin2  9  +  cos2  ( 


\Fr 


+ 

dFr 


sin2  6>  +  cos2  9  \  dFe 


36 


dFz 
IF 


1         1  dFe  8F, 
dr       r         r  d9  dz 


( 


\ 


r  h  Fr  H  rF, 

3r  3(9  3zV 


as  desired. 


In  spherical  coordinates,  the  story  for  the  gradient,  divergence,  and  curl  is 
more  complicated  algebraically,  although  the  ideas  behind  the  proof  are  essentially 


3.4  I  Exercises  235 


the  same.  We  state  the  relevant  results  and  leave  to  you  the  rather  tedious  task  of 
verifying  them. 


THEOREM  4.6  Let  /:  X  C  R3  ->  R  and  F:  Y  C  R3  ->  R3  be  differentiable 
scalar  and  vector  fields,  respectively.  Then  the  following  formulas  hold: 


v/  = 

1  df 
 e 

p  d<p 

1 

p  sin<p 

deee; 

V-F  = 

1   3  , 

pL  dp 

%)  + 

1  3 
p  sin<p  3^> 

(sin<p  Fy)  + 

1 

eP 

p  sin  <p  e# 

V  x  F  = 

d/dp 

d/dcp 

3/36* 

p2  sin  <p 

p  sin  Fg 

1 


3Fe 


p  sin<p  dO 


(8) 
(9) 

(10) 


3.4  Exercises 


Calculate  the  divergence  of  the  vector  fields  given  in  Exer- 
cises 1-6. 


1.  F 

3.  F 

4.  F 

5.  F 

6.  F 


x2i  +  y2] 


2.  F 


y2i  +  x2} 


(x  +  y)i  +  (y  +  z)j  +  (jc  +  z)k 
Z  cos  (er )  i  +  xy/z2  +  1  j  +  e2v  sin  3jc  k 

Xjei  +  2x|e2  H  h  «x2e„ 

xiei  +  2xie2  +  •  •  •  +  nx\e„ 
Find  the  curl  of  the  vector  fields  given  in  Exercises  7-11. 

7.  F  =  x2  i  —  xey  j  +  2xyz  k 

8.  F  =  xi  +  yj  +  zk 

9.  F  =  (x  +  yz)i  +  (y  +  xz)}  +  (z  +  xy)k 

1 0.  F  =  (cos  yz  —  x)i  +  (cos  xz  —  y)j  +  (cos  xy  —  z)k 

11.  F  =  y2zi  +  exyzj+x2yk 

12.  (a)  Consider  again  the  vector  field  in  Exercise  8  and 

its  curl.  Sketch  the  vector  field  and  use  your  pic- 
ture to  explain  geometrically  why  the  curl  is  as  you 
calculated. 

(b)  Use  geometry  to  determine  V  x  F,  where  F  = 
(xi  +  yj  +  zk) 

y/x2  +  y2  +  z2 ' 

(c)  For  F  as  in  part  (b),  verify  your  intuition  by  explic- 
itly computing  V  x  F. 

13.  Can  you  tell  in  what  portions  of  R2,  the  vector  fields 
shown  in  Figures  3.43-3.46  have  positive  divergence? 
Negative  divergence? 


\  \  \  \  i  i  t 

\  \  \  \  i  i  i 

w  \  u  I  I 

v  x  \  \  \  \  I 


/  /  I   1    I    i  \ 

/  /  I  I  I  I  I 

////Ml 


1  1  t  t  t  /  / 
III//// 
I  /  /  t  f  /  / 

I    /    /    /   /  S  S 


I  \  \  \  \  \  N 
I  I  \  \  \  \  \ 
\  M  \  \  \  \ 


Figure  3.43  Vector  field  for  Exercise  13(a). 


s  \  \  \  \ 


i  i  /  /  / 


S  S  /    /    /     /  \     \    \    \   *s  ^  ^ 


Figure  3.44  Vector  field  for  Exercise  13(b). 


Chapter  3  |  Vector-Valued  Functions 


Figure  3.45  Vector  field  for  Exercise  13(c). 


////II  I 
////ill 
///iii  i 
✓  ✓///// 


X  \  \  \  \  \  \ 
S  \  \  \  \  \  | 
X    \    \     \     \     \  ( 

\\  \  \  \  u 


M  \  \  V  \  s 

\  \  \  \  \  \  \ 

\     V  \  \  \  V  V 

\     S  \  \  \  \  N 


III//// 
I  I  I  I  /  /  / 
III//// 
III//// 


Figure  3.46  Vector  field  for  Exercise  13(d). 

14.  Check  that  if  f(x,  y,z)  =  x2  sin  y  +  y2  cosz,  then 

V  x  (V/)  =  0. 

15.  Check  that  if  F(x,  y,  z)  =  xyzi  —  ez  cos xj  +  xy2z3K 
then 

V  •  (V  x  F)  =  0. 

16.  Prove  Theorem  4.4. 

In  Exercises  1 7-20,  let  r  =  x  i  +  y  j  +  zkand  let  r  denote  ||r||. 
Verify  the  following: 

17.  Vr"  =nr"-2r 

18.  V(lnr)=  ^ 

ri 

19.  V  •  (r"r)  =  (n  +  3)r" 

20.  V  x  (rnr)  =  0 

In  Exercises  21—25,  establish  the  given  identities.  (You  may 
assume  that  any  functions  and  vector  fields  are  appropriately 
differentiable.) 

21.  V-(F  +  G)  =  V-F  +  V-G 


22.  V  x  (F  +  G)  =  V  x  F  +  V  x  G 

23.  V.(/F)  =  /V.F  +  F.V/ 

24.  V  x  (/F)  =  /V  x  F  +  V/  x  F 

25.  V-(FxG)  =  G-  VxF-F-VxG 

26.  Prove  formulas  (3)  and  (5)  of  Theorem  4.5. 

27.  Establish  the  formula  for  the  gradient  of  a  function  in 
spherical  coordinates  given  in  Theorem  4.6. 

28.  The  Laplacian  operator,  denoted  V2,  is  the  second- 
order  partial  differential  operator  defined  by 

,     s2     s2  a2 

V2  =  1  1  . 

dx2      3y2  dz2 

(a)  Explain  why  it  makes  sense  to  think  of  V2  as  V  •  V . 

(b)  Show  that  if  /  and  g  are  functions  of  class  C2 ,  then 

V2(/g)  =  fV2g  +  gV2f  +  2(V/  •  Vg). 

(c)  Show  that 

V-(/Vg-gV/)  =  /V2g-gV2/. 

29.  Show  that  V  •  (/V/)  =  ||  V/||2  +  /V2/. 

30.  Show  that  V  x  (V  x  F)  =  V(V  •  F)  -  V2F.  (Here 
V2F  means  to  take  the  Laplacian  of  each  component 
function  of  F.) 

Let  X  be  an  open  set  in  R",  F:XC  R"  R"  a  vector  field 
on  X,  and  a  €  X.  If  y  is  any  unit  vector  in  R",  we  define  the 
directional  derivative  of  'F  at  a  in  the  direction  of  v,  denoted 
DvF(a),  by 

1 

DvF(a)  =  lim  -(F(a  +  hv)  -  F(a)), 

h^O  h 

provided  that  the  limit  exists.  Exercises  31-34  involve  direc- 
tional derivatives  of  vector  fields. 

31.  (a)  In  analogy  with  the  directional  derivative  of  a 

scalar-valued  function  defined  in  §2.6,  show  that 


DvF(a) 


dt 


F(a  + 1  v) 


[=0 


(b)  Use  the  result  of  part  (a)  and  the  chain  rule  to  show 
that,  if  F  is  differentiable  at  a,  then 

DvF(a)  =  DF(a)v, 

where  v  is  interpreted  to  be  an  n  x  1  matrix.  (Note 
that  this  result  makes  it  straightforward  to  calculate 
directional  derivatives  of  vector  fields.) 

32.  Show  that  the  directional  derivative  of  a  vector  field 
F  is  the  vector  whose  components  are  the  directional 


Miscellaneous  Exercises  for  Chapter  3  237 


derivatives  of  the  component  functions  F\, . . . ,  F„  of 
F,  that  is,  that 

DvF(a)  =  (DyFii*),  DvF2(a), DvFn(a)). 

33.  LetF  =  vzi  +  xzj +xyk.FindD(i_j+k)/V3F(3,2,  1). 
(Hint:  See  Exercise  31.) 


34.  Let  F  =  x  i  +  y  j  +  z  k.  Show  that  DvF(a)  =  v  for  any 
point  a  €  R3  and  any  unit  vector  veR3.  More  gener- 
ally, if  F  =  (x\,  X2,  ■  ■  ■ ,  xn),  a  =  (fli,  ai, . . . ,  an),  and 
v  =  (t>i,  V2  vn),  show  that  _DvF(a)  =  v. 


True/False  Exercises  for  Chapter  3 


1 .  If  a  path  x  remains  a  constant  distance  from  the  origin, 
then  the  velocity  of  x  is  perpendicular  to  x. 

2.  If  a  path  is  parametrized  by  arclength,  then  its  velocity 
vector  is  constant. 

3.  If  a  path  is  parametrized  by  arclength,  then  its  velocity 
and  acceleration  are  orthogonal. 

'  ■x(f)|[  =  ||x'(f)[|. 


4. 


dt 
d 


5.  — (x  x  y) 

dtK  3 


6.  K 


dT 

dt 


dy  dx 

xx  h  y  x  — . 

dt     J  dt 


7.  |T| 


dB 

ds 


8.  The  curvature  k  is  always  nonnegative. 

9.  The  torsion  r  is  always  nonnegative. 

10.  N 


dT 

ds 


11.  If  a  path  x  has  zero  curvature,  then  its  acceleration  is 
always  parallel  to  its  velocity. 

12.  If  a  path  x  has  a  constant  binormal  vector  B,  then  r  =0. 


|a(r)[|2 


14.  grad  /  is  a  scalar  field. 

15.  div  F  is  a  vector  field. 

16.  curl  F  is  a  vector  field. 


1 7.  grad(div  F)  is  a  vector  field. 

18.  div(curl(grad  /))  is  a  vector  field. 

1 9.  grad/  x  div  F  is  a  vector  field. 

20.  The  path  x(f)  =  (2  cos  t,  4  sinr,  t)  is  a  flow  line  of  the 

y 

vector  field  F(x ,y,  z)  =  —  —  i  +  2x  j  +  z  k. 

21.  The  path  x(?)  =  (e'  cosf ,  e'(cos  t  +  sinf),  e'  sin?)  is  a 
flow  line  of  the  vector  field  F(jc,  y,  z)  =  (x  —  z)i  + 
2jc  j  +  y  k. 

22.  The  vector  field  F  =  2xy  cos  z  i  —  y2  cos  z  j  +  exy  k  is 
incompressible. 

23.  The  vector  field  F  =  2xy  cos  z  i  —  y2  cos  z  j  +  exy  k  is 
irrotational. 

24.  V  x  (V/)  =  0  for  all  functions  /:  R3  — »  R. 

25.  If  V  •  F  =  0  and  V  x  F  =  0,  then  F  =  0. 

26.  V  •  (F  x  G)  =  F  •  (V  x  G)  +  G  •  (V  x  F). 

27.  If  F  =  curl  G,  then  F  is  solenoidal. 

28.  The  vector  field  F  =  2x  sin  y  cos  z  i  +  x2  cos  y  cos  z 
j  +  x2  sin  y  sin  z  k  is  the  gradient  of  a  function  /  of 
class  C2. 

29.  There  is  a  vector  field  F  of  class  C2  on  R3  such  that 
V  x  F  =  x  cos2  y  i  +  3y  j  —  xyz2  k. 

30.  If  F  and  G  are  gradient  fields,  then  F  x  G  is  incom- 
pressible. 


Miscellaneous  Exercises  for  Chapter  3 


1 .  Figure  3.47  shows  the  plots  of  six  paths  x  in  the  plane. 
Match  each  parametric  description  with  the  correct 
graph. 

(a)  x(/)  =  (s'mlt,  sin 3i) 

(b)  x(f)  =  (t  +  sin5f,  t2  +  cos  6?) 

(c)  x(r)  =  (f2  +  l,f3-?) 

(d)  x(f)  =  (2?  +  sin4f,  t  -  sin5f) 

(e)  x(t)  =  {t  -t2,t3  -t) 

(f)  x(t)  =  (sin(?  +  sin 30,  cosf) 


2.  Figure  3.48  shows  the  plots  of  six  paths  x  in  R3 .  Match 
each  parametric  description  with  the  correct  graph. 

(a)  x(f)  =  (t  +  cos3f,  t2  +  sin5f,  sin  At) 

(b)  x(f)  =  (2  cos3 1,  3  sin3 1,  cos2f) 

(c)  x(0  =  (15  cos  /,  23  sin  t ,  At) 

(d)  x(0  =  (cos3r,  cos5f,  sin4f) 

(e)  x(f)  =  (2f  cosf,  2f  sin?,  4f) 

(f)  x(f)  =  (r2  +  l,?3-f,?4-r2) 


238       Chapter  3  |  Vector-Valued  Functions 


Miscellaneous  Exercises  for  Chapter  3  239 


3.  Suppose  that  x  is  a  C2  path  with  nonzero  velocity.  Show 
that  x  has  constant  speed  if  and  only  if  its  velocity  and 
acceleration  vectors  are  always  perpendicular  to  one 
another. 

4.  You  are  at  Vertigo  Amusement  Park  riding  the  new 
Vector  roller  coaster.  The  path  of  your  car  is  given  by 


x(?)  =  (x(t),  y{t)),  where 


x(f) 


t/60        1Tt     t/60   ■  Jtt 

e  '    cos  —  ,e  '    sm  — , 
30  30 

2r(10  -  t)(t  -  90)2 


80  + 


106 


where  /  =  0  corresponds  to  the  beginning  of  your 
three-minute  ride,  measured  in  seconds,  and  spatial 
dimensions  are  measured  in  feet.  It  is  a  calm  day,  but 
after  90  sec  of  your  ride  your  glasses  suddenly  fly  off 
your  face. 

(a)  Neglecting  the  effect  of  gravity,  where  will  your 
glasses  be  2  sec  later? 

(b)  What  if  gravity  is  taken  into  account? 

5.  Show  that  the  curve  traced  parametrically  by 

1 


x(f)  =  ^cos(f  -  1),  t  -  1, 

istangenttothesurface.ii;3  +  y3  +  z3 
t  =  1. 


xyz  =  0  when 


6.  Gregor,  the  cockroach,  is  on  the  edge  of  a  Ferris  wheel 
that  is  rotating  at  a  rate  of  2  rev/min  (counterclock- 
wise as  you  observe  him).  Gregor  is  crawling  along 
a  spoke  toward  the  center  of  the  wheel  at  a  rate  of 
3  in/min. 

(a)  Using  polar  coordinates  with  the  center  of  the 
wheel  as  origin,  assume  that  Gregor  starts  (at  time 
t  =  0)  at  the  point  r  =  20  ft,  6  =  0.  Give  paramet- 
ric equations  for  Gregor's  polar  coordinates  r  and 
6  at  time  t  (in  minutes). 

(b)  Give  parametric  equations  for  Gregor's  Cartesian 
coordinates  at  time  t. 

(c)  Determine  the  distance  Gregor  has  traveled  once 
he  reaches  the  center  of  the  wheel.  Express  your 
answer  as  an  integral  and  evaluate  it  numerically. 

If  you  have  used  a  drawing  program  on  a  computer,  you  have 
probably  worked  with  a  curve  known  as  a  Bezier  curve.1  Such 
a  curve  is  defined  parametrically  by  using  several  control 
points  in  the  plane  to  shape  the  curve.  In  Exercises  7-12, 
we  discuss  various  aspects  of  quadratic  Bezier  curves.  These 
curves  are  defined  by  using  three  fixed  control  points  {x\ ,  y\), 
(x2,  yi),  and(xi,  yi)  and  a  nonnegative  constant  w.  The  Bezier 
curve  defined  by  this  information  is  given  by  x:  [0,  1]  — >  R2, 


X(t): 

y(t) : 


(1  -  tfx\  +  2wt(l  -  t)x2  +  t2X} 
(\-tf  +  2wt(\-t)  +  t2 

(1  -tfyi  +2wtQ_  -r)y2  +  f2y3 
(1  -t)2  +  2wt{\  -t)  +  t2 


0  <  t  <  1. 


(1) 

^  7.  Let  the  control  points  be  (1,0),  (0,  1),  and  (1,  1). 
Use  a  computer  to  graph  the  Bezier  curve  for  w  = 
0,  1/2,  1,  2,  5.  What  happens  as  w  increases? 

8.  Repeat  Exercise  7  for  the  control  points  (—1,  —  1), 
(1,3),  and  (4,  1). 

9.  (a)  Show  that  the  Bezier  curve  given  by  the  paramet- 

ric equations  in  ( 1 )  has  (x\ ,  y\ )  as  initial  point  and 
(x3,  V3)  as  terminal  point. 

(b)  Show  that  x(i)  lies  on  the  line  segment  joining 
(*2,  yi)  to  the  midpoint  of  the  line  segment  joining 
(x\,y\)  to  (x3,  y3). 

10.  In  general  the  control  points  (xi,y{),  (x2,y2),  and 
(*3 ,  )?3 )  will  form  a  triangle,  known  as  the  control  poly- 
gon for  the  curve.  Assume  in  this  problem  that  w  >  0. 
By  calculating  x'(0)  and  x'(l),  show  that  the  tangent 
lines  to  the  curve  at  x(0)  and  x(l)  intersect  at  (X2,  V2). 
Hence,  the  control  triangle  has  two  of  its  sides  tangent 
to  the  curve. 

11.  In  this  problem,  you  will  establish  the  geometric  sig- 
nificance of  the  constant  w  appearing  in  the  equations 
in(l). 

(a)  Calculate  the  distance  a  between  x(  j )  and  (x2 ,  ^2). 

(b)  Calculate  the  distance  b  between  x(^)  and  the 
midpoint  of  the  line  segment  joining  (xi,  y\)  and 

C*3,  yj)- 

(c)  Show  that  w  =  b/a.  By  part  (b)  of  Exercise  9, 
x(|)  divides  the  line  segment  joining  (X2,  3^)  to 
the  midpoint  of  the  line  segment  joining  (x\,  y\) 
to  (*3,  V3)  into  two  pieces,  and  w  represents  the 
ratio  of  the  lengths  of  the  two  pieces. 

12.  Determine  the  Bezier  parametrization  for  the  portion 
of  the  parabola  y  =  x2  between  the  points  (—2,  4)  and 
(2,  4)  as  follows: 

(a)  Two  of  the  three  control  points  must  be  (—2,  4)  and 
(2,  4).  Find  the  third  control  point  using  the  result 
of  Exercise  10. 

(b)  Using  part  (a)  and  Exercise  9,  we  must  have  that 
x(i)  lies  on  the  y-axis  and,  hence,  at  the  point 


2  P.  Bezier  was  an  automobile  design  engineer  for  Renault.  See  D.  Cox,  J.  Little,  and  D.  O'Shea,  Ide- 
als, Varieties,  and  Algorithms:  An  Introduction  to  Computational  Algebraic  Geometry  and  Commutative 
Algebra,  3rd  ed.  (Springer- Verlag,  New  York,  2007),  pp.  28-29.  Exercises  7-1 1  adapted  with  permission. 


240       Chapter  3  |  Vector- Valued  Functions 


(0,  0).  Use  the  result  of  Exercise  11  to  determine 
the  constant  w. 

(c)  Now  write  the  Bezier  parametrization.  You  should 
be  able  to  check  that  your  answer  is  correct. 

1 3.  Let  x:  (0,  7t)  -*■  R2  be  the  path  given  by 

x(?)  =  (sin?,  cos?  +  In  tan  |)  , 

where  ?  is  the  angle  that  the  y-axis  makes  with  the 
vector  x(?).  The  image  of  x  is  called  the  tractrix.  (See 
Figure  3.49.) 

(a)  Show  that  x  has  nonzero  speed  except  when  ?  = 
n/2. 

(b)  Show  that  the  length  of  the  segment  of  the  tangent 
to  the  tractrix  between  the  point  of  tangency  and 
the  y-axis  is  always  equal  to  1.  This  means  that 
the  image  curve  has  the  following  description:  Let 
a  horse  pull  a  heavy  load  by  a  rope  of  length  1 . 


of  class  C2.  Use  equation  (17)  in  §3.2  to  derive  the 
curvature  formula 


Figure  3.49  The  tractrix 
of  Exercise  13. 

Suppose  that  the  horse  initially  is  at  (0,  0),  the  load 
at  (1,  0),  and  let  the  horse  walk  along  the  v-axis. 
The  load  follows  the  image  of  the  tractrix. 

14.  Another  way  to  parametrize  the  tractrix  path  given  in 
Exercise  13  is 

y:  (-oo,  O)-*-  R2, 


where  y(r) 


O-P 


dp 


(a)  Show  that  y  satisfies  the  property  described  in  part 
(b)  of  Exercise  13. 

(b)  In  fact,  y  is  actually  a  reparametrization  of  part  of 
the  path  x  of  Exercise  13.  Without  proving  this  fact 
in  detail,  indicate  what  portion  of  the  image  of  x 
the  image  of  y  covers. 

15.  Suppose  that  a  plane  curve  is  given  in  polar  coordi- 
nates by  the  equation  r  =  f{6),  where  /  is  a  function 


K(9) 


'  +  2r': 


(r2  _|_  r/2)3/2 


(Hint:  First  give  parametric  equations  for  the  curve  in 
Cartesian  coordinates  using  9  as  the  parameter.) 

16.  Use  the  result  of  Exercise  15  to  calculate  the  curvature 
of  the  lemniscate  r2  =  cos  26. 

Let  x:  /  — >  R2  be  a  path  of  class  C2  that  is  not  a  straight  line 
and  such  that  x'(?)  ^  0.  Choose  some  to  €  I  and  let 

y(?)  =  x(?)  -  s(f)T(f), 

where  s(t)  =  f'  ||x'(r)||  dr  is  the  arclength  function  and  T  is 
the  unit  tangent  vector.  The  path  y:  /  —>  R2  is  called  the  invo- 
lute of  x.  Exercises  1 7—19  concern  involutes  of  paths. 

17.  (a)  Calculate  the  involute  of  the  circular  path  of  radius 

a,  that  is,  x(?)  =  (a  cos  ?,  a  sin  ?).  (Take  ?o  to  be  0.) 

(b)  Let  a  =  1  and  use  a  computer  to  graph  the  path  x 
and  the  involute  path  y  on  the  same  set  of  axes. 

1 8.  Show  that  the  unit  tangent  vector  to  the  involute  at  ? 
is  the  opposite  of  the  unit  normal  vector  N(?)  to  the 
original  path  x.  (Hint:  Use  the  Frenet-Serret  formulas 
and  the  fact  that  a  plane  curve  has  torsion  equal  to  zero 
everywhere.) 

19.  Show  that  the  involute  y  of  the  path  x  is  formed  by 
unwinding  a  taut  string  that  has  been  wrapped  around 
x  as  follows: 

(a)  Show  that  the  distance  in  R2  between  a  point  x(?) 
on  the  original  path  and  the  corresponding  point 
y(?)  on  the  involute  is  equal  to  the  distance  traveled 
from  x(?o)  to  x(?)  along  the  underlying  curve  of  x. 

(b)  Show  that  the  distance  between  a  point  x(?)  on  the 
path  and  the  corresponding  point  y(?)  on  the  in- 
volute is  equal  to  the  distance  from  x(?)  to  y(?) 
measured  along  the  tangent  emanating  from  x(?). 
Then  finish  the  argument. 

Let  x:  /  — >  R2  be  a  path  of  class  C2  that  is  not  a  straight  line 
and  such  that  x'(?)  /  0.  Let 

e(?)  =  x(?)  +  -N(?). 

K 

This  is  the  path  traced  by  the  center  of  the  osculating  circle  of 
the  path  x.  The  quantity  p  =  \/k  is  the  radius  of  the  osculat- 
ing circle  and  is  called  the  radius  of  curvature  of  the  path  x. 
The  path  e  is  called  the  evolute  of  the  path  x.  Exercises  20—25 
involve  evolutes  of  paths. 

20.  Letx(?)  =  (?,  ?2)beaparabolicpath.(SeeFigure3.50.) 

(a)  Find  the  unit  tangent  vector  T,  the  unit  normal 
vector  N,  and  the  curvature  k  as  functions  of  ?. 

(b)  Calculate  the  evolute  of  x. 


Miscellaneous  Exercises  for  Chapter  3  241 


(c)  Use  a  computer  to  plot  x(f)  and  e(?)  on  the  same 
set  of  axes. 


y 


Figure  3.50  The  parabola  and  its 
osculating  circle  at  a  point.  The  centers 
of  the  osculating  circles  at  all  points  of 
the  parabola  trace  the  evolute  of  the 
parabola  as  described  in  Exercise  20. 


21 .  Show  that  the  evolute  of  a  circular  path  is  a  point. 

22.  (a)  Use  a  computer  algebra  system  to  calculate  the  for- 

mula for  the  evolute  of  the  elliptical  path  x(f)  = 
(a  cost,  b  sinr). 

(b)  Use  a  computer  to  plot  x(f )  and  the  evolute  e(f )  on 
the  same  set  of  axes  for  various  values  of  the  con- 
stants a  and  b.  What  happens  to  the  evolute  when 
a  becomes  close  in  value  to  bl 

23.  Use  a  computer  algebra  system  to  calculate  the  formula 
for  the  evolute  of  the  cycloid  x(f )  =  (at  —  a  sin  t,  a  — 
a  cost).  What  do  you  find? 

24.  Use  a  computer  algebra  system  to  calculate  the  formula 
for  the  evolute  of  the  cardioid  x(f)  =  (2a  cosf(l  + 
a  cos  t),  2a  sin  t (1  +  a  cos  t )). 

25.  Assuming  ic'(t)  ^  0,  show  that  the  unit  tangent  vector 
to  the  evolute  e(f )  is  parallel  to  the  unit  normal  vector 
N(f )  to  the  original  path  x(f). 

26.  Suppose  that  a  C1  path  x(f )  is  such  that  both  its  veloc- 
ity and  acceleration  are  unit  vectors  for  all  t .  Show  that 
k  =  1  for  all  t. 

27.  Consider  the  plane  curve  parametrized  by 


x(s)  =  /   cos  g(t)dt,     y(s)  =  I    sin  g(t)dt, 
Jo  Jo 

where  g  is  a  differentiable  function. 

(a)  Show  that  the  parameter  s  is  the  arclength  param- 
eter. 

(b)  Calculate  the  curvature  k(s). 


(c)  Use  part  (b)  to  explain  how  you  can  create  a 
parametrized  plane  curve  with  any  specified  con- 
tinuous, nonnegative  curvature  function  k(s). 

(d)  Give  a  set  of  parametric  equations  for  a  curve 
whose  curvature  k(s)  =  \s\.  (Your  answer  should 
involve  integrals.) 

(e)  Use  a  computer  to  graph  the  curve  you  found  in 
part  (d),  known  as  a  clothoid  or  a  spiral  of  Cornu. 
(Note:  The  integrals  involved  are  known  as  Fres- 
nel  integrals  and  arise  in  the  study  of  optics.  You 
must  evaluate  these  integrals  numerically  in  order 
to  graph  the  curve.) 

28.  Suppose  that  x  is  a  C3  path  in  R3  with  torsion  r  always 
equal  to  0. 

(a)  Explain  why  x  must  have  a  constant  binormal  vec- 
tor (i.e.,  one  whose  direction  must  remain  fixed  for 
all  f). 

(b)  Suppose  we  have  chosen  coordinates  so  that  x(0)  = 
0  and  that  v(0)  and  a(0)  lie  in  the  .ty-plane  (i.e., 
have  no  k-component).  Then  what  must  the  binor- 
mal vector  B  be? 

(c)  Using  the  coordinate  assumptions  in  part  (b),  show 
that  x(f)  must  lie  in  the  jc;y-plane  for  all  t .  (Hint: 
Begin  by  explaining  why  v(f)  •  k  =  a(r)  •  k  =  Ofor 
all  t.  Then  show  that  if 

x(t)  =  x(t)i  +  y(t)j  +  z(t)k, 

we  must  have  z(t )  =  0  for  all  t .) 

(d)  Now  explain  how  we  may  conclude  that  curves 
with  zero  torsion  must  lie  in  a  plane. 

29.  Suppose  that  x  is  a  C3  path  in  R3,  parametrized  by  arc- 
length,  with  k  /  0.  Suppose  that  the  image  of  x  lies  in 
the  xy -plane. 

(a)  Explain  why  x  must  have  a  constant  binormal 
vector. 

(b)  Show  that  the  torsion  r  must  always  be  zero. 
Note  that  there  is  really  nothing  special  about  the  im- 
age of  x  lying  in  the  Jfj-plane,  so  that  this  exercise, 
combined  with  the  results  of  Exercise  28,  shows  that 
the  image  of  x  is  a  plane  curve  if  and  only  if  r  is  always 
zero  and  if  and  only  if  B  is  a  constant  vector. 

30.  In  Example  7  of  §3.2  we  saw  that  if  x  is  a  straight-line 
path,  then  x  has  zero  curvature.  Demonstrate  the  con- 
verse; that  is,  if  x  is  a  C2  path  parametrized  by  arc- 
length  s  and  has  zero  curvature  for  all  s,  then  x  traces 
a  straight  line. 

31 .  A  large  piece  of  cylindrical  metal  pipe  is  to  be  manu- 
factured to  include  a  strake,  which  is  a  spiraling  strip 
of  metal  that  offers  structural  support  for  the  pipe.  (See 
Figure  3.51.)  The  pieces  of  the  strake  are  to  be  made 
from  flat  pieces  of  flexible  metal  whose  curved  sides 
are  arcs  of  circles  as  shown  in  Figure  3.52.  Assume  that 


242       Chapter  3  |  Vector- Valued  Functions 


the  pipe  has  a  radius  of  a  ft  and  that  the  strake  makes 
one  complete  revolution  around  the  pipe  every  h  ft.3 


Figure  3.51  A  cylindrical         Figure  3.52  A  section  of 
pipe  with  strake  attached.  the  strake.  (See  Exercise  3 1 .) 

(a)  In  terms  of  a  and  h,  what  should  the  inner  radius  r 
be  so  that  the  strake  will  fit  snugly  against  the  pipe? 

(b)  Suppose  a  =  3  ft  and  h  =  25  ft.  What  is  r? 

Suppose  that  x:  /  — >  R3  is  a  path  of  class  C3  parametrized  by 
arclength.  Then  the  unit  tangent  vector  T(s)  defines  a  vector- 
valued function  T:  /  —>  R3  that  may  also  be  considered  to  be  a 
path  (although  not  necessarily  one  parametrized  by  arclength, 
nor  necessarily  one  with  nonvanishing  velocity).  Since  T  is  a 
unit  vector,  the  image  of  the  path  T  must  lie  on  a  sphere  of 
radius  1  centered  at  the  origin.  This  image  curve  is  called  the 
tangent  spherical  image  of  x.  Likewise,  we  may  consider  the 
functions  defined  by  the  normal  and  binormal  vectors  N  and  B 
to  give  paths  called,  respectively,  the  normal  spherical  image 
and  binormal  spherical  image  of  x.  Exercises  32-35  concern 
these  notions. 

32.  Find  the  tangent  spherical  image,  normal  spherical 
image,  and  binormal  spherical  image  of  the  circular 
helix  x(/)  =  (a  cosf ,  a  sinf ,  bt).  (Note:  The  path  x  is 
not  parametrized  by  arclength.) 

33.  Suppose  that  x  is  parametrized  by  arclength.  Show 
that  x  is  a  straight-line  path  if  and  only  if  its  tangent 
spherical  image  is  a  constant  path.  (See  Example  7  of 
§3.2  and  Exercise  30.) 

34.  Suppose  that  x  is  parametrized  by  arclength.  Show  that 
the  image  of  x  lies  in  a  plane  if  and  only  if  its  binormal 
spherical  image  is  constant.  (See  Exercises  28  and  29.) 

35.  Suppose  that  x  is  parametrized  by  arclength.  Show 
that  the  normal  spherical  image  of  x  can  never  be 
constant. 

36.  In  this  problem,  we  will  find  expressions  for  velocity 
and  acceleration  in  cylindrical  coordinates.  We  begin 


with  the  expression 

x(t)  =  x(t)i  +  y(t)j  +  z(t)k 

for  the  path  in  Cartesian  coordinates. 

(a)  Recall  that  the  standard  basis  vectors  for  cylindri- 
cal coordinates  are 

e,-  =  cosf?  i  +  s'md  j, 
ee  =  —  s'md  i  +  cos#  j, 
e,  =  k. 

Use  the  facts  that  x  =  r  cos  0  and  y  =  r  sin  0  to 
show  that  we  may  write  x(r )  as 

x(t)  =  r(t)er+z(t)ez. 

(b)  Use  the  definitions  of  er,  ee,  and  ez  just  given  and 
the  chain  rule  to  find  der/dt,  deg/dt,  and  de./dt 
in  terms  of  er ,  eg ,  and  ez . 

(c)  Now  use  the  product  rule  to  give  expressions  for  v 
and  a  in  terms  of  the  standard  basis  for  cylindrical 
coordinates. 

37.  Suppose  that  the  path 

x(f)  =  (sin2f,  Vlcoslt,  sin2f  —  2) 

describes  the  position  of  the  Starship  Inertia  at  time  t . 

(a)  Lt.  Commander  Agnes  notices  that  the  ship  is  trac- 
ing a  closed  loop.  What  is  the  length  of  this  loop? 

(b)  Ensign  Egbert  reports  that  the  Inertia's  path  is 
actually  a  flow  line  of  the  Martian  vector  field 
F(jc,  y,  z)  =  yi  —  2x\  +  yk,  but  he  omitted  a  con- 
stant factor  when  he  entered  this  information  in 
his  log.  Help  him  set  things  right  by  finding  the 
correct  vector  field. 

38.  Suppose  that  the  temperature  at  points  inside  a  room  is 
given  by  a  differentiable  function  T(x,  y,  z).  Livinia, 
the  housefly  (who  is  recovering  from  a  head  cold),  is  in 
the  room  and  desires  to  warm  up  as  rapidly  as  possible. 

(a)  Show  that  Livinia 's  path  x(?)  must  be  a  flow  line 
of  kVT,  where  &  is  a  positive  constant. 

(b)  If  T(x,  y,  z)  =  x2  —  2y2  +  3z2  and  Livinia  is  ini- 
tially at  the  point  (2,3,-1),  describe  her  path 
explicitly. 

39.  Let  F  =  u(x,  y)i  —  v(x,  y)j  be  an  incompressible, 
irrotational  vector  field  of  class  C2. 

(a)  Show  that  the  functions  u  and  v  (which  deter- 
mine the  component  functions  of  F)  satisfy  the 
Cauchy-Riemann  equations 

9m      3d  du  dv 

—  =  — ,    and    —  =  . 

dx       dv  dy  ox 


3  See  F.  Morgan,  Riemannian  Geometry:  A  Beginner's  Guide,  2nd  ed.  (A  K  Peters,  Wellesley,  1998), 
pp.  7-10.  Figures  3.51  and  3.52  adapted  with  permission. 


Miscellaneous  Exercises  for  Chapter  3  243 


(b)  Show  that  u  and  v  are  harmonic,  that  is,  that 


(Also  see  §1.4  concerning  the  notion  of  torque.)  Show 
that 


d2u  d2U 
9^  +  9^2 


40.  Suppose  that  a  particle  of  mass  m  travels  along  a  path 
x  according  to  Newton's  second  law  F  =  ma,  where 
F  is  a  gradient  vector  field.  If  the  particle  is  also  con- 
strained to  lie  on  an  equipotential  surface  of  F,  show 
that  then  it  must  have  constant  speed. 

41 .  Let  a  particle  of  mass  m  travel  along  a  differentiable 
path  x  in  a  Newtonian  vector  field  F  (i.e.,  one  that 
satisfies  Newton's  second  law  F  =  ma,  where  a  is  the 
acceleration  of  x).  We  define  the  angular  momen- 
tum 1(f)  of  the  particle  to  be  the  cross  product  of  the 
position  vector  and  the  linear  momentum  mv,  that  is, 

l(r)  =  x(f)  x  m\(t). 

(Here  v  denotes  the  velocity  of  x.)  The  torque  about 
the  origin  of  the  coordinate  system  due  to  the  force  F 
is  the  cross  product  of  position  and  force: 

M(f )  =  x(f)  x  F(f)  =  x(f)  x  ma(f). 


dt 


M. 


Thus,  we  see  that  the  rate  of  change  of  angular  mo- 
mentum is  equal  to  the  torque  imparted  to  the  particle 
by  the  vector  field  F. 

42.  Consider  the  situation  in  Exercise  41  and  suppose  that 
F  is  a  central  force  (i.e.,  a  force  that  always  points 
directly  toward  or  away  from  the  origin).  Show  that  in 
this  case  the  angular  momentum  is  conserved,  that  is, 
that  it  must  remain  constant. 

43.  Can  the  vector  field 

F  =  (ex  cos  y  +  e~ x  sin  z)  i  —  ex  sin  y  j  +  e~x  cos  z  k 

be  the  gradient  of  a  function  f(x,  y,  z)  of  class  C2? 
Why  or  why  not? 

44.  Can  the  vector  field 

F  =  *(y2  +  \)i  +  (yex  -  ez) j  +  x2ez  k 

be  the  curl  of  another  vector  field  G(x,  y,  z)  of  class 
C2?  Why  or  why  not? 


4 Maxima  and  Minima 
in  Several  Variables 


4.1  Differentials  and  Taylor's 
Theorem 

4.2  Extrema  of  Functions 

4.3  Lagrange  Multipliers 

4.4  Some  Applications  of 
Extrema 

True/False  Exercises  for 
Chapter  4 

Miscellaneous  Exercises 
for  Chapter  4 


4.1    Differentials  and  Taylor's  Theorem 

Among  all  classes  of  functions  of  one  or  several  variables,  polynomials  are  without 
a  doubt  the  nicest  in  that  they  are  continuous  and  differentiable  everywhere  and 
display  intricate  and  interesting  behavior.  Our  goal  in  this  section  is  to  provide 
a  means  of  approximating  any  scalar-valued  function  by  a  polynomial  of  given 
degree,  known  as  the  Taylor  polynomial.  Because  of  the  relative  ease  with  which 
one  can  calculate  with  them,  Taylor  polynomials  are  useful  for  work  in  computer 
graphics  and  computer-aided  design,  to  name  just  two  areas. 

Taylor's  Theorem  in  One  Variable:  A  Review 

Suppose  you  have  a  function  /:lcR->R  that  is  differentiable  at  a  point  a  in 
X.  Then  the  equation  for  the  tangent  line  gives  the  best  linear  approximation  for 
/  near  a.  That  is,  when  we  define  p\  by 

p\{x)  =  f(a)  +  f'(a)(x  —  a),     we  have    pi(x)  &  fix)  if  x  %  a. 

(See  Figure  4.1.)  As  explained  in  §2.3,  the  phrase  "best  linear  approximation" 
means  that  if  we  take  R\(x,  a)  to  be  f(x)  —  p\(x),  then 


lim 


Ri(x,  a) 


0. 


Note  that,  in  particular,  we  have  pi(a)  =  f(a)  and  p[(a)  =  f'(a). 

Generally,  tangent  lines  approximate  graphs  of  functions  only  over  very  small 
neighborhoods  containing  the  point  of  tangency.  For  a  better  approximation,  we 
might  try  to  fit  a  parabola  that  hugs  the  function's  graph  more  closely  as  in 
Figure  4.2.  In  this  case,  we  want  p2  to  be  the  quadratic  function  such  that 

p2(a)  =  f(a),     p'2(a)  =  f'(a),     and    p'2\a)  =  f"(a). 

The  only  quadratic  polynomial  that  satisfies  these  three  conditions  is 

P2(x)  =  f(a)  +  f(a){x  -a)+  ^(x  -  a)2. 


It  can  be  proved  that,  if  /  is  of  class  C2,  then 


fix)  =  p2(x)+  R2(x,  a), 


4.1  |  Differentials  and  Taylor's  Theorem  245 


where 


Figure  4.3 

fix)  =  In* 


y =  2- 

Approximations  to 


Riix,  a) 
lim  =  0. 

*-»-a  ix  —  a)1 


EXAMPLE  1    If  fix)  =  In  x,  then,  for  a  =  1 ,  we  have 

/(l)  =  lnl=0, 
1 
I 

1 

12 


/'(I) 
/"(I)  = 


1. 


1. 


Hence, 


piix)  =  0  +  1(* 
p2(x)  =  0+  1(jc 


1)  =  *-  1, 

1)  -  I(*  -  l)2  = 


2x 


The  approximating  polynomials  p\  and  p2  are  shown  in  Figure  4.3.  ♦ 

There  is  no  reason  to  stop  with  quadratic  polynomials.  Suppose  we  want  to 
approximate  /  by  a  polynomial  of  degree  k,  where  k  is  a  positive  integer. 
Analogous  to  the  work  above,  we  require  that  pk  and  its  first  k  derivatives  agree 
with  /  and  its  first  k  derivatives  at  the  point  a.  Thus,  we  demand  that 


Pk(a)  = 

f{a), 

Pk(a)  = 

f'(a), 

P'l(a)  = 

pf  («)  =  /w(a). 

Given  these  requirements,  we  have  only  one  choice  for  pk,  stated  in  the  following 
theorem: 


-  f(k)( 


THEOREM  1.1  (Taylor's  theorem  in  one  variable)  Let  X  be  open  in 
R  and  suppose  /:XcR^R  is  differentiable  up  to  (at  least)  order  k. 


246       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


Given  a  e  X,  let 

f"(a)  f(kXa) 
pk(x)  =  f{a)  +  f'(a)(x  -a)+  J-^-(x  -  af  +  •  •  •  +  J—^-{x  -  af.  (1) 

2  k\ 

Then 

f(x)  =  Pk(x)+  Rk{x,a), 
where  the  remainder  term  Rk  is  such  that  Rk(x,  a)/(x  —  af  -¥  0  as  x  —*  a. 


Figure  4.4  The  graphs  of 

(1)  y  =x  -a, 

(2)  y  =  (x  —  a)2,  and 

(3)  y  =  (x-  af. 

Note  how  much  more  closely  the 
graph  of  (3)  hugs  the  jc-axis  than 
thatof(l)or(2). 


The  polynomial  defined  by  formula  (1)  is  called  the  &th-order  Taylor  poly- 
nomial of  /  at  a.  The  essence  of  Taylor's  theorem  is  this:  For  x  near  a,  the  Taylor 
polynomial  pk  approximates  /  in  the  sense  that  the  error  Rk  involved  in  making 
this  approximation  tends  to  zero  even  faster  than  (x  —  af  does.  When  k  is  large, 
this  is  very  fast  indeed,  as  we  see  graphically  in  Figure  4.4. 

EXAMPLE  2   Consider  In x  with  a  =  1  again.  We  calculate 
/(l)  =  lnl  =  0, 

/'(l)  =  y  =  l> 


/"(I) 


1 

12 


1* 


i^l  =  (-l)k+\k  -  1)!. 


Therefore, 

Pk(x)  =  (x 


lf  +  -3(x 


If 


+ 


(-1) 


<X  ~  I  f 


Taylor's  theorem  as  stated  in  Theorem  1.1  says  nothing  explicit  about  the 
remainder  term  Rk.  However,  it  is  possible  to  establish  the  following  derivative 
form  for  the  remainder: 


PROPOSITION  1 .2  If  /  is  of  class  Ck+\  then  there  exists  some  number  z  be- 
tween a  and  x  such  that 


Rk(x,  a)  =  f(  +  )(Z\x  -  af+\ 
(Jfc  +  1)! 


(2) 


In  practice,  formula  (2)  is  quite  useful  for  estimating  the  error  involved  with  a 
Taylor  polynomial  approximation.  Both  Theorem  1 . 1  (under  the  slightly  stronger 
hypothesis  that  /  is  of  class  Ck+ 1 )  and  Proposition  1 .2  are  proved  in  the  addendum 
to  this  section. 


EXAMPLE  3  The  fifth-order  Taylor  polynomial  of  f(x)=  cos  x  about  x  =  tt/2 
is  i  «       1  < 

/         7T\        1  /         7T\3  1     /  7T\5 

P5(x)  =  -\x-2)  +  6\x-2)  -mV'V  ■ 


4.1  |  Differentials  and  Taylor's  Theorem  247 


(You  should  verify  this  calculation.)  According  to  formula  (2),  the  difference 
between  p5  and  cos  x  is 


R, 


(      n\        f(6\z)  (         11  \6  COSZ  /  7t\( 

\X'2)  =  ~^\X~2)  =~~6T\X~2) 


2/        6!     V       2/  6!    V  2. 

where  z  is  some  number  between  n/2  and  x.  Since  |  cosjcI  is  never  larger  than  1, 
we  have 

^       7T\6  ^(jC-TT/2)6 


*5  (*,  |) 


cos  z 


6! 


Thus,  for  x  in  the  interval  [0,  it],  we  have 

(tt  -  7T/2)6 


R5  (x.  |) 


71' 


720 


46,080 


720 


0.0209. 


In  other  words,  the  use  of  the  polynomial  ps  above  in  place  of  cosx  will  be 
accurate  to  at  least  0.0209  throughout  the  interval  [0,  n].  ♦ 

Taylor's  Theorem  in  Several  Variables: 

The  First-order  Formula  


For  the  moment,  suppose  that  /:  X  c  R2  — >  R  is  a  function  of  two  variables, 
where  X  is  open  in  R2  and  of  class  Cl.  Then  near  the  point  (a,  b)  e  X,  the  best 
linear  approximation  to  /  is  provided  by  the  equation  giving  the  tangent  plane  at 
(a,  b,  f(a,  b)).  That  is, 

f(x,  y)  »  pi(x,  v), 

where 

pi(x,  y)  =  f(a,  b)  +  fx(a,  b)(x  -a)  +  fy(a,  b)(y  -  b). 

Note  that  the  linear  polynomial  p\  has  the  property  that 

pi(a,  b)  =  f(a,  b); 

dpi  df 

-7—  {a,b)=  —(a,b), 

ox  ox 

dpi  df 
-5—  (a,  b)  =  —(a,  b). 
oy  oy 

Such  an  approximation  is  shown  in  Figure  4.5. 

To  generalize  this  situation  to  the  case  of  a  function  /:  X  c  R"  — >  R  of 
class  C1,  we  naturally  use  the  equation  for  the  tangent  hyperplane.  That  is,  if 


z=f(x,y) 


Figure  4.5  The  graph  of  z  =  f(x,  y)  and 
z  =  pi(x,  y). 


248       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


a  =  (ai,  a2, . . . ,  a„)  e  X,  then 

f(Xl,X2,  ...,xn)&  pi{xi,x2,  ...,x„), 

where 

Pi(xi,       xn)  =  /(a)  +  fxi(a)(xi  -  ai)  +  /X2(a)(x2  -  a2) 

+  r-  fx„(a)(xn  ~  an)- 

Of  course,  the  formula  for  p\  can  be  written  more  compactly  using  either 
E -notation  or,  better  still,  matrices: 

n 

Pi(jci,  . . . ,  xn)  =  /(a)  +  £  /*(»)(*  -  a,)  =  /(a)  +  D/(a)(x  -  a).  (3) 

;  =  1 

EXAMPLE  4   Let  /(xj,  X2,  X3,  X4)  =  xi  +  2x2  +  3x3  +  4x4  +  X1X2X3X4.  Then 

3/      1  3/  „ 

 =  l  +  X2X3X4,   =  2  +  X1X3X4, 

3xj  3x2 

3/      „  3/ 

- —  =  3+xix2x4,  - — =4  +  X!X2x3. 

dX3  dX4 

At  a  =  0  =  (0,  0,  0,  0),  we  have 

^(0)  =  l,        ^(0)  =  2,        |^(0)  =  3,        |^(0)  =  4. 

dx\  6x2  0x3  dx4 

Thus, 

piixi,  x2,  x3,  x4)  =  0  +  l(xi  -  0)  +  2(x2  -  0)  +  3(x3  -  0)  +  4(x4  -  0) 

=  X\  +  2X2  +  3X3  +  4X4. 

Note  that  p\  contains  precisely  the  linear  terms  of  the  original  function  /.  On  the 
other  hand,  if  a  =  (1 ,  2,  3,  4),  then 

-Ml,  2,  3, 4)  =  25,       —^-(1,2,  3,4)  =  14, 
dxi  dx2 

df  df 

-Ml,  2,  3,  4)=  11,       -Ml,  2,  3,  4)  =  10, 

dX3  dX4 

so  that,  in  this  case, 

pi(xi,X2,x3,x4)  =  54  +  25(xi  -  1)+  14(x2  -  2)  +  ll(x3  -3)+  10(x4  -  4). 

♦ 

The  relevant  theorem  regarding  the  first-order  Taylor  polynomial  is  just  a  re- 
statement of  the  definition  of  differentiability.  However,  since  we  plan  to  consider 
higher-order  Taylor  polynomials,  we  state  the  theorem  explicitly. 

THEOREM  1 .3  (First-order  Taylor's  formula  in  several  variables)  Let 
X  be  open  in  R"  and  suppose  that  /:XcR"->R  is  differentiate  at  the  point 
a  in  X.  Let 

pi(x)  =  /(a)  +  D/(aXx-a)-  (4) 

Then 

/(x)  =  pi(x)  +  *i(x,  a), 
where  Ri(x,  a)/||x  —  a||  — >•  0  as  x  —>  a. 


4.1  |  Differentials  and  Taylor's  Theorem  249 

Note  that  we  may  also  express  the  first-order  Taylor  polynomial  using  the  gradient. 
In  place  of  (4),  we  would  have 

Pi(x)  =  /(a)  +  V/(a).(x-a). 

Differentials   

Before  we  explore  higher-order  versions  of  Taylor's  theorem  in  several  variables, 
we  consider  the  linear  (or  first-order)  approximation  in  further  detail. 
Let  h  =  x  —  a.  Then  formula  (3)  becomes 

Pl(x)  =  /(a)  +  fl/(a)h  =  m  +  ■  (5) 

i=i  dx> 

We  focus  on  the  sum  appearing  in  formula  (5)  and  summarize  its  salient 
features  as  follows: 


DEFINITION  1 .4  Let  /  :  X  C  R"  ->  R  and  let  ael  The  incremental 
change  of  /,  denoted  Af,  is 

A/  =  /(a  +  h)-/(a). 

The  total  differential  of  /,  denoted  df(a,  h),  is 

df  df  df 

dX\  0X2  dxn 

The  significance  of  the  differential  is  that  for  h  «  0, 

Af  «  df. 
(We  have  abbreviated  df(a,  h)  by  J /.) 


Sometimes  hj  is  replaced  by  the  expression  Ax,  or  dx{  to  emphasize  that  it 
represents  a  change  in  the  z'th  independent  variable,  in  which  case  we  write 

At        9/  A  A  V  A 

df  =  —dxl  +  —dx2-\  h  —  dxn. 

dx\  0X2  6xn 

(We've  suppressed  the  evaluation  of  the  partial  derivatives  at  a,  as  is  customary.) 
EXAMPLE  5    Suppose  f(x,  y,  z)  =  sin(xyz)  +  cos(xyz).  Then 

Af      V  j    .  &f  ,  df 
df  =  —  dx  +  —  dy  +  —  dz 
6x  6y  6z 

=  yz[cos(xyz)  —  sm(xyz)]dx  +  xz[cos(xyz)  —  sin(xyz)]dy 

+  xy[cos(xvz)  —  sin(xyz)]dz 
=  (cos(xyz)  —  sin(xyz))(yz  dx  +  xz  dy  +  xy  dz)-  ^ 

The  geometry  of  the  differential  arises,  naturally  enough,  from  tangent  lines 
and  planes.  (See  Figures  4.6  and  4.7.)  In  particular,  the  incremental  change  Af 
measures  the  change  in  the  height  of  the  graph  of  /  when  moving  from  a  to  a  +  h; 
the  differential  change  d f  measures  the  corresponding  change  in  the  height  of 


250       Chapter  4  [  Maxima  and  Minima  in  Several  Variables 


Figure  4.6  The  incremental 
change  A/  equals  the  change  in 
y  -coordinate  of  the  graph  of 
y  =  f(x)  as  the  x-coordinate  of  a 
point  changes  from  a  to  a  +  dx. 
The  differential  df  equals  the 
change  in  y -coordinate  of  the 
graph  of  the  tangent  line  at  a  (i.e., 
the  graph  of  y  =  p\(x)). 


Figure  4.7  The  incremental 
change  Af  equals  the  change  in 
^-coordinate  of  the  graph  of 
z  =  f(x,  y)  as  a  point  in  R2 
changes  from  a  =  (a,  b)  to 
a  +  h  =  (a  +  h,  b  +  k).  The 
differential  df  equals  the  change  in 
z-coordinate  of  the  graph  of  the 
tangent  plane  at  (a,  b). 


the  graph  of  the  (hyper)plane  tangent  to  the  graph  at  a.  When  h  is  small  (i.e., 
when  a  +  h  is  close  to  a),  the  differential  df  approximates  the  increment  Af  and 
it  is  often  easier  from  a  technical  standpoint  to  work  with  the  differential. 

EXAMPLE  6  Let  /(*,  y)  =  x  -  y  +  2x2  +  xy2.  Then  for  (a,  b)  =  (2,-1), 
we  have  that  the  increment  is 

Af  =  f(2  +  Ax,  -1  +  A.y)  -  f(2,  -1) 

=  2  +  Ax  -  (-1  +  Ay)  +  2(2  +  Ax)2  +  (2  +  Ax)(-1  +  A.y)2  -  13 
=  lOAx  -  5 Ay  +  2(Ax)2  -  2Ax Ay  +  2(A^)2  +  Ax(Ay)2. 

On  the  other  hand, 

df((2,  -1),  (Ax,  Ay))  =  fx(2,  -I) Ax  +  fy(2,  -I) Ay 

=  (1  +  4x  +  .v2)|(2,-i)Ax  +  (-1  +  2x>>)|(2,-i)A.y 
=  lOAx  -  5A;y. 

We  see  that  df  consists  of  exactly  the  terms  of  Af  that  are  linear  in  Ax  and  Ay 
(i.e.,  appear  to  first  power  only).  This  will  always  be  the  case,  of  course,  since 
that  is  the  nature  of  the  first-order  Taylor  approximation.  Use  of  the  differential 
approximation  is  often  sufficient  in  practice,  for  when  Ax  and  Ay  are  small,  higher 
powers  of  them  will  be  small  enough  to  make  virtually  negligible  contributions 
to  Af.  For  example,  if  Ax  and  Ay  are  both  0.01,  then 

df  =  (0.1  -0.05)  =  0.05 

and 

Af  =  (0.1  -  0.05)  +  0.0002  -  0.0002  +  0.0002  +  0.000001 
=  0.05  +  0.000201  =  0.050201. 

Thus,  the  values  of  df  and  Af  are  the  same  to  three  decimal  places.  ♦ 


4.1  |  Differentials  and  Taylor's  Theorem  251 


EXAMPLE  7  A  wooden  rectangular  block  is  to  be  manufactured  with  dimen- 
sions 3  in  x  4  in  x  6  in.  Suppose  that  the  possible  errors  in  measuring  each  di- 
mension of  the  block  are  the  same.  We  use  differentials  to  estimate  how  accurately 
we  must  measure  the  dimensions  so  that  the  resulting  calculated  error  in  volume 
is  no  more  than  0.1  in3 . 

Let  the  dimensions  of  the  block  be  denoted  by  x      3  in),  y      4  in),  and  z 
6  in).  Then  the  volume  of  the  block  is 

V=xyz    and    V  «  3  ■  4  •  6  =  72  in3. 

The  error  in  calculated  volume  is  A  V,  which  is  approximated  by  the  total  differ- 
ential dV.  Thus, 

A V  w  d V  =  Vx(3,  4,  6)Ax  +  Vy(3,  4,  6)Ay  +  V,(3,  4,  6)AZ 
=  24Ax  +  18Ay  +  12Az. 

If  the  error  in  measuring  each  dimension  is  e,  then  we  have  Ax  =  Ay  =  Az  =  e. 
Therefore, 

dV  =  24Ax  +  18Ay  +  12Az  =  24e  +  18e  +  12e  =  54e. 
To  ensure  (approximately)  that  |  A  V\  <  0.1,  we  demand 

\dV\  =  |54e|  <  0.1. 

Hence, 

0.1 

|e|  <  —  =  0.0019  m. 
"  54 

So  the  measurements  in  each  dimension  must  be  accurate  to  within  0.0019  in.  ♦ 


Figure  4.8  Which  would 
you  buy? 


EXAMPLE  8  The  formula  for  the  volume  of  a  cylinder  of  radius  r  and  height 
h  is  V(r,  h)  =  jtr2h.  If  the  dimensions  are  changed  by  small  amounts  Ar  and 
Ah,  then  the  resulting  change  A  V  in  volume  is  approximated  by  the  differential 
change  d  V.  That  is, 


AV  ^dV  = 


dV         dV  , 

—  Ar  H  Ah  =  InrhAr  +  nr  Ah. 

dr  dh 


Suppose  the  cylinder  is  actually  a  beer  can,  so  that  it  has  approximate  dimensions 
of  r  =  1  in  and  h  =  5  in.  Then 

dV  =  jr(10Ar  +  Ah). 

This  statement  shows  that,  for  these  particular  values  of  r  and  h,  the  volume 
is  approximately  10  times  more  sensitive  to  changes  in  radius  than  changes  in 
height.  That  is,  if  the  radius  is  changed  by  an  amount  e,  then  the  height  must  be 
changed  by  roughly  lOe  to  keep  the  volume  constant  (i.e.,  to  make  AV  zero). 
We  use  the  word  "approximate"  because  our  analysis  arises  from  considering  the 
differential  change  d  V  rather  than  the  actual  incremental  change  A  V. 

This  beer  can  example  has  real  application  to  product  marketing  strategies. 
Because  the  volume  is  so  much  more  sensitive  to  changes  in  radius  than  height, 
it  is  possible  to  make  a  can  appear  to  be  larger  than  standard  by  decreasing  its 
radius  slightly  (little  enough  so  as  to  be  hardly  noticeable)  and  increasing  the 
height  so  no  change  in  volume  results.  (See  Figure  4.8.)  This  sensitivity  analysis 
shows  that  even  a  tiny  decrease  in  radius  can  force  an  appreciable  compensating 
increase  in  height.  The  result  can  be  quite  striking,  and  these  ideas  apparently 


252       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


have  been  adopted  by  at  least  one  brewery.  Indeed  this  is  how  the  author  came  to 
fully  appreciate  differentials  and  sensitivity  analysis.1  ♦ 

Taylor's  Theorem  in  Several  Variables: 

The  Second-order  Formula  

Suppose  /:  X  c  R2  —>  R  is  a  C2  function  of  two  variables.  Then  we  know  that 
the  tangent  plane  gives  rise  to  a  linear  approximation  p\  of  /  near  a  given  point 
(a,  b)  of  X.  We  can  improve  on  this  result  by  looking  for  the  quadric  surface  that 
best  approximates  the  graph  of  z  =  fix,  y)  near  (a,  b,  f(a,  b)).  See  Figure  4.9 
for  an  illustration.  That  is,  we  search  for  a  degree  2  polynomial  pi(x,  y)  =  Ax2  + 
Bxy  +  Cy2  +  Dx  +  Ey  +  F  such  that,  for  (x,  y)     (a,  b), 

f(x,  y)  «  p2(x,  y). 


Quadric 
surface 


Z=f(x,y) 


Figure  4.9  The  tangent  plane  and  quadric 
surface. 


Analogous  to  the  linear  approximation  pi,  it  is  reasonable  to  require  that  p2  and 
all  of  its  first-  and  second-order  partial  derivatives  agree  with  those  of  /  at  the 
point  (a,  b).  That  is,  we  demand 


dpi 
dx 

d2P2 
dx2 


(a,  b)  = 


(a,b) 


p2(a,  b) 

— ( 

dx 

d2f 
— — ( 
dx2 

d2p2 


(a,b), 
(a,  b), 

2  (a,  b) 


f(a,b) 
dp2 


dy 

d2P2 
dxdy 

d2f 


(a,  b)  = 


(a,b)  = 


df 

dy 
d2f 


dxdy 


{a,  b), 
(a,b), 


(6) 


dy 


dy: 


(a,b). 


After  some  algebra,  we  see  that  the  only  second-degree  polynomial  meeting  these 
requirements  is 

P2(x,  y)  =  f(a,  b)  +  fx(a,  b)(x  -a)  +  fy(a,  b)(y  -  b) 


+  yxx(a,  b)(x  -  a)2  +  fxy(a,  b)(x  -  a)(y  -  b) 
+  \fyy(a,b)(y-b)2. 


(7) 


1  See  S.  J.  Colley,  The  College  Mathematics  Journal,  25  (1994),  no.  3,  226-227.  Art  reproduced  with 
permission  from  the  Mathematical  Association  of  America. 


4.1  |  Differentials  and  Taylor's  Theorem  253 


How  does  formula  (7)  generalize  to  functions  of  n  variables?  We  need  to  begin 
by  demanding  conditions  analogous  to  those  in  (6)  for  a  function  /:  X  cR"->R. 
For  a  =  (a\,  a2,  . . . ,  an)  £  X,  these  conditions  are 

p2(a)  =  /(a), 

-^(a)=-^(a),         i  =  l,2,...,n,  (8) 

dxi  dxi 

d2p2  d2f 

(a)=  — ^-(a),    ij  =  1,2,..., n. 


dxidxj  dxidxj 

If  you  do  some  algebra  (which  we  omit),  you  will  find  that  the  only  polynomial 
of  degree  2  that  satisfies  the  conditions  in  (8)  is 

n  \  H 

P2(x)  =  /(a)  +       A,(a)(*<  "  ad  +2  E  fxi*M(Xi  ~  ai)(xJ  -  (9) 

1=1  i.j=l 

(Note  that  the  second  sum  appearing  in  (9)  is  a  double  sum  consisting  of  n2  terms.) 
To  check  that  everything  is  consistent  when  n  =  2,  we  have 

p2(xi,x2)  =  f{a\,a2)  +  fxM^'  ai)(xi  -  fli)  +  fx2(ai,a2)(x2  -  a2) 

+  \  \_Sx^xSfl\^i){x\  -  aif  +  fXlX2(ai,  a2){x\  -  a{)(x2  -  a2) 
+  fx2Xl(ai,a2)(x2  -  a2){x\  -  d\)  +  fXlX2{a\,  a2)(x2  -  a2)2]  . 

When  /'  is  a  C2  function,  the  two  mixed  partials  are  the  same,  so  this  formula 
agrees  with  formula  (7). 

EXAMPLE  9   Let  /(*,  y,  z)  =  ex+y+z  and  let  a  =  (a,  b,  c)  =  (0,  0,  0).  Then 
/(0,0,  0)  =  e°=  1, 

fx(0,  0,  0)  =  /,(0,  0,  0)  =  fz(0,  0,  0)  =  e°  =  1, 
fxx(Q,  0,  0)  =  fxy(0,  0,  0)  =  fxt(0,  0,  0)  =  fyy(0,  0,  0) 
=  fyz(0,  0,  0)  = /„((),  0,  0)  =  e°  =  1. 

Thus, 

P2(x,  y,z)  =  l  +  l(x  -  0)  +  l(y  -  0)  +  l(z  -  0) 

+  \  [l(x  -  0)2  +  2  ■  l(x  -  0)(y  -  0)  +  2  ■  l(x  -  0)(z  -  0) 
+  l(y  -  0)2  +  2  •  l(y  -  0)(z  -  0)  +  l(z  -  0)2] 

=  1  +  x  +  y  +  z  +  k2  +  xy  +  xz  +       +  yz  +  \z2 


=  1  +  (x  +  y  +  z)  +  Ux  +  y  +  zf 


2 


We  have  made  use  of  the  fact  that,  since  /  is  of  class  C  ,  a  term  like 

fxy(0,  0,  0)(x  -  0)(y  -  0)    is  equal  to    fyx(0,  0,  0)(y  -  0)(x  -  0). 


Now  we  state  the  second-order  version  of  Taylor's  theorem  precisely. 


254       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


THEOREM  1 .5  (Second-order  Taylor's  formula)  Let  X  be  open  in  R" ,  and 
suppose  that  /:XcR"->R  is  of  class  C2.  Let 

n  j  n 

P2OO  =  /(a)  +         fxi(&Xxi  ~  ad  +  j  Y  fxixMXXi  ~  a^XJ  ~  ai)" 
i=l  Z  <,J=1 

Then 

/(x)  =  p2(x)  +  fl2(x,  a), 
where  |/?2|/||x  -  all2  ->  0  as  x  ->•  a. 

A  version  of  Theorem  1.5,  under  the  stronger  assumption  that  /  is  of  class  C3,  is 
established  in  the  addendum  to  this  section. 

EXAMPLE  10   Let  f(x,  y)  =  cosx  cosy  and  (a,  b)  =  (0,  0).  Then 

/(0,0)  =  1; 

fx(0,  0)  =  -  sin  x  cosy  I  (0,0)  =  0,        fy(0,  0)  =  -  cosx  siny|(0,0)  =  0; 
fxx(0,  0)  =  -  cosxcosy|(0  0)  =  -1, 

fxy(Q,  0)  =  sinx  siny|(0  0)  =  0, 

/yy(0,0)  =  -  cosxcosy|(00)  =  -1. 

Hence, 

f(x,  y)  «  p2(x,  y)  =  1  +  \{-\  ■  x2  -  1  •  y2)  =  1  -  \x2  -  \y2 . 

We  can  also  solve  this  problem  another  way  since  /  is  a  product  of  two  functions. 
We  can  multiply  the  two  Taylor  polynomials: 

Pi(x,  y)  =  (Taylor  polynomial  for  cos  x)  •  (Taylor  polynomial  for  cos  y) 
=  (1  -  \x2)  (1  -  \y2) 

=  1  —  \x2  —  \y2  up  to  terms  of  degree  2. 

This  method  is  justified  by  noting  that  if  qi  is  the  Taylor  polynomial  for  cosine 
and  Ro  is  the  corresponding  remainder  term,  then 

cosx  cosy  =  [q2(x)  +  R2(x,  0)][q2(y)  +  R2(y,  0)] 

=  qi(x)q2(y)  +  q2(y)R2(x,  0)  +  q2(x)R2(y,  0)  +  R2(x,  0)R2(y,  0) 

=  q2(x)q2(y)  +  other  stuff, 

where  (other  stuff)/||(x,  y)||2  —>  0  as  (x,  y)  (0,  0),  since  both  R2{x,  0)  and 
R2(y,  0)  do.  ♦ 


The  Hessian 

Recall  that  the  formula  for  the  first-order  Taylor  polynomial  p\  was  written  quite 
concisely  in  formula  (5)  by  using  vector  and  matrix  notation.  It  turns  out  that  it 
is  possible  to  do  something  similar  for  the  second-order  polynomial  p2. 


4.1  |  Differentials  and  Taylor's  Theorem  255 


DEFINITION  1 .6    The  Hessian  of  a  function  /:XcR"->R  is  the  matrix 

\x/,Vir\cf1  /  t  1"Vi  ptitr\7  ic        f  1  r)  v 
YVllUov  £  /  111  C11L1  y  la  u     /  /  (JA 

j  OJij .  1  Hal  la, 

fx,X!  fx,X2 

fx\x„ 

Hf  = 

fx2X\  fx2x2 

fx2x„ 

_  fx„xi  fx„x2 

fxnx„ 

The  term  "Hessian"  comes  from  Ludwig  Otto  Hesse,  the  mathematician  who 
first  introduced  it,  not  from  the  German  mercenaries  who  fought  in  the  American 
revolution. 

Now  let's  look  again  at  the  formula  for  p2  in  Theorem  1.5: 

n  j  n 

P2(x)  =  /(a)  +  £  fXi(*)hi  +  fx,*  (»)/<;/',. 

(We  have  let  h  =  (hi, . . . ,  hn)  =  x  —  a.)  This  can  be  written  as 


p2«  =  /(a)+[/,1(a)  /T2(a)  ■■■  A„(a)] 


&2 


+ 


1  r 


/i2   ■■•  h„ 


Thus,  we  see  that 


/*2»(a)  A2*2(a)  •••  A2X„(a) 
_/*„*,  (a)  /x„.v2(a)  •••  A,,,, (a)  J  _ 


h2 


p2(x)  =  /(a)  +  D/(a)h  +  ^hr///(a)h.  (10) 


(Remember  that  hT  is  the  transpose  of  the  w  x  1  matrix  h.) 

EXAMPLE  11  (Example  10  revisited)  For  f(x,  y)  =  cosx  cos  y,  a  =  (0,  0), 

we  have 

Df(x,  y)  =  [—  sinx  cosy   —  cosx  sin y] 

and 


Hf(x,y) 


cos  x  cos  y  sin  x  sin  y 
sin  x  sin  y   —  cos  x  cos  y 


256       Chapter  4  [  Maxima  and  Minima  in  Several  Variables 


Hence, 


P2(x,  y)  =  /(O,  0)  +  Df(0,  0)h  +  lhTHf(0,  0)h 


l+[ 0  0] 


+  \[h,  h2] 


'  -1      0  " 

"  hi  ' 

0  -1 

=  1 


l-h2 

2  1 


X-h2 
2n2- 


Once  we  recall  that  h  =  (hi,  h2)  =  (x  —  0,  y  —  0)  =  (x,  y),  we  see  that  this  result 
checks  with  our  work  in  Example  10,  just  as  it  should.  ♦ 


Higher-order  Taylor  Polynomials   

So  far  we  have  said  nothing  about  Taylor  polynomials  of  degree  greater  than  2 
in  the  case  of  functions  of  several  variables.  The  main  reasons  for  this  are  (i)  the 
general  formula  is  quite  complicated  and  has  no  compact  matrix  reformulation 
analogous  to  (10)  and  (ii)  we  will  have  little  need  for  such  formulas  in  this  text. 
Nonetheless,  if  your  curiosity  cannot  be  denied  here  is  the  third-order  Taylor 
polynomial  for  a  function  /:  X  c  R"  —>  R  of  class  C3  near  aeX: 


P3(x)  =  /(a)  +      fxi(*)(xi  ~  ad  +  ~  ^  -  atXxj  -  aj) 

i=l  i,j=\ 

1  " 

i,j,k=l 


(The  relevant  theorem  regarding  pi  is  that  f(x)  —  pj,(x)  +  Rj,(x,  a),  where 
|/?3(x,  a)|/||x  —  a  || 3  ->  0  as  x  ->  a.)  If  you  must  know  even  more,  the  /cth-order 
Taylor  polynomial  is 


i=l  Z  i,y=l 

1  " 

H  h  jti    XI    A  (a^Xi'i  ~  fl,'i) ' ' '  ^  ~  a<*  )• 

l'l,...,H  =  l 


Formulas  for  Remainder  Terms  (optional) 

Under  slightly  stricter  hypotheses  than  those  appearing  in  Theorems  1.3  and 
1.5,  integral  formulas  for  the  remainder  terms  may  be  derived  as  follows.  Set 
h  =  x  —  a.  If  /  is  of  class  C2,  then 

n      p  1 

Ri(x,  a)  =  V  /  (1  -  /  )/,  ,, (a  +  th)h,hj  dt 

=  [  [hT#/(a  +  fh)h](l -t)dt. 
Jo 


4.1  |  Differentials  and  Taylor's  Theorem  257 


If  /  is  of  class  C3,  then 

v  r<i2 


^     (l  -  o2 

R2(x,  a)  =  2_,   /  — ^ — ./',,«,,  (a  +  th)hihjhk  dt, 


and  if  /  is  of  class  C*+1,  then 

i?,(x,a)=     J]     /  — — — fxiyxi2-xit+1(a  +  t\L)hhhi2  ■  ■  ■  hh+l  dt. 
U  '*+i=i  Jo 

Although  explicit,  these  formulas  are  not  very  useful  in  practice.  By  artful  appli- 
cation of  Taylor's  formula  for  a  single  variable,  we  can  arrive  at  derivative  versions 
of  these  remainder  terms  (known  as  Lagrange's  form  of  the  remainder)  that  are 
similar  to  those  in  the  one-variable  case. 


Lagrange's  form  of  the  remainder.  If  /  is  of  class  C2,  then  in  Theorem  1 .3 
the  remainder  Ri  is 


1  " 

#i(x,  a)  =  -  J]  fXiXj(z)hihj 


2  UM 

for  a  suitable  point  z  in  the  domain  of  /  on  the  line  segment  joining  a  and 
x  =  a  +  h.  Similarly,  if  /  is  of  class  C3,  then  the  remainder  R2  in  Theorem 
1.5  is 


1  - 

tf2(x,  a)=  —  fXiXjXk{z)hihjhk 


31 

ij,k=l 

for  a  suitable  point  z  on  the  line  segment  joining  a  and  x  =  a  +  h.  More 
generally,  if  /  is  of  class  Ck+l,  then  the  remainder  Rk  is 

*^X' a)  =  TkTiv  fw-*** V)KK  ■  ■  ■  K+i 

K  <i,-,i*+i=1 

for  a  suitable  point  z  on  the  line  segment  joining  a  and  x  =  a  +  h. 


The  remainder  formulas  above  are  established  in  the  addendum  to  this  section. 

EXAMPLE  12   For  fix,  y)  =  cos  x  cos  y,  we  have 

2 

X]  fxiXjXk(z)hihjhk 


\R2(x,y,0,  0)|  =  i 


2 


< 

tj,*=i 

since  all  partial  derivatives  of  /  will  be  a  product  of  sines  and  cosines  and,  hence, 
no  larger  than  1  in  magnitude.  Expanding  the  sum,  we  get 

\R2(x,y,0,  0)|  <  i(|/Jl|3  +  3/i?|/!2|  +  3|/i1|^  +  |/z2|3). 

If  both  \hi  \  and  \  fi2\  are  no  more  than,  say,  0.1,  then 

\R2(x,y,0,  0)|  <  i  (8-  (O.l)3)  =  0.0013. 


Chapter  4  i  Maxima  and  Minima  in  Several  Variables 


y 


z 


0.1 


Z  =  cos  x  cos  y 


-0.1 


0.1 


X 


y 


-0.1 


X 


Figure  4.11  The  graph  of  f(x,  y)  = 
cos  x  cos  y  and  its  Taylor  polynomial 


Figure  4.10  The  polynomial  p2 
approximates  /  to  within  0.0013 
on  the  square  shown.  (See 
Example  12.) 


Pi(x,  y)  =  1  —  \x2  —  \y2  over  the  square 
{(Jt.y)  1-1  <*<!,-!  <?<!}. 


So  throughout  the  square  of  side  0.2  centered  at  the  origin  and  shown  in  Fig- 
ure 4.10,  the  second-order  Taylor  polynomial  is  accurate  to  at  least  0.0013  (i.e.,  to 
two  decimal  places)  as  an  approximation  of  f(x,  y)  =  cosx  cosy.  In  Figure  4. 11, 
we  show  the  graph  of  f(x,  y)  —  cos  x  cos  y  over  the  square  domain  {(x ,  y)  |  —  1  < 
x  <  1 ,  —  1  <  y  <  1}  together  with  the  graph  of  its  second-order  Taylor  polyno- 
mial p2(x,  y)  —  1  —  \x2  —  \y2  (calculated  in  Example  10).  Note  how  closely  the 
surfaces  coincide  near  the  point  (0,  0,  1),  just  as  the  analysis  above  indicates.  ♦ 

Addendum:  Proofs  of  Theorem  1.1,  Proposition  1.2, 
and  Theorem  1.5 

Below  we  establish  some  of  the  fundamental  results  used  in  this  section.  We  begin 
by  proving  Theorem  1.1,  Taylor's  theorem  for  function  of  a  single  variable,  and 
Proposition  1 .2  regarding  the  remainder  term  in  Theorem  1.1.  We  then  use  these 
results  to  "bootstrap"  a  proof  of  the  multivariable  result  of  Theorem  1.5  and  to 
derive  Lagrange's  formula  for  the  remainder  term  appearing  in  it. 

Proof  of  Theorem  1 .1  We  prove  the  result  under  the  stronger  assumption  that  / 
is  of  class  Ck+l  rather  than  assuming  that  /  is  only  differentiate  up  to  order  k. 
(This  distinction  matters  little  in  practice.) 
By  the  fundamental  theorem  of  calculus, 


We  evaluate  the  integral  on  the  right  side  of  ( 1 1 )  by  means  of  integration  by  parts. 
Recall  that  the  relevant  formula  is 


We  use  this  formula  with  u  =  f'(t)  and  v  =  x  —  t  so  that  dv  =  —dt.  (Note  that 
in  the  right  side  of  (1 1),  x  plays  the  role  of  a  constant.)  We  obtain 


(11) 


u  dv  =  uv  — 


(12) 


4.1  |  Differentials  and  Taylor's  Theorem  259 


Combining  (11)  and  (12),  we  have 

fix)  =  f(a)  +  f'(a)(x  -a)  +  \\x  -  t)f"(t)dt.  (13) 
Thus,  we  have  shown,  when  /  is  differentiable  up  to  (at  least)  second  order,  that 
Ri(x,a)=  f  (x  -  t)f"(t)dt. 

J  a 

This  provides  an  integral  formula  for  the  remainder  in  formula  ( 1 )  of  Theorem  1 . 1 
when  k  =  1,  although  we  have  not  yet  established  that  R\(x,a)/(x  —  a)  — >  0  as 
x  — >  a. 

To  obtain  the  second-order  formula,  the  case  k  =  2  of  (1),  we  focus  on 
Ri(x,  a)  =  f*(x  —  t)f"{t)dt  and  integrate  by  parts  again,  this  time  with  u  = 
f"(t)  and  v  =  (x  -  tf/2,  so  that  dv  =  -(x  -  t)dt.  We  obtain 


f\x 

J  a 


t)f"(t)dt  = 


f"(t)(x  -  tf 


f"(a)(x  -  a)2 


a  Ja 

f 


f"'(t)dt 


(x  -  tf 


f"'(t)dt. 


Hence  (13)  becomes 

f"(a) 

f(x)  =  f(a)  +  f'(a)(x  -a)+  J—^-{x 


2    r*  (x-t) 

a)  +L  — 


Therefore,  we  have  shown,  when  /  is  differentiable  up  to  (at  least)  third  order, 
that 

-Jlf"\t)dt. 


R2(x,a)=  f 

J  a 


We  can  continue  to  argue  in  this  manner  or  use  mathematical  induction  to  show 
that  formula  (1)  holds  in  general  with 

Rk(x,a)  =  j"^t^\t)dt,  (14) 

assuming  that  /  is  differentiable  up  to  order  (at  least)  k+\. 

It  remains  to  see  that  Rk(x,  a)/(x  —  a)k  — >  0  as  x  — >  a.  In  formula  (14)  we 
are  only  considering  t  between  a  and  x,  so  that  \x  —  t  \  <  \x  —  a\.  Moreover,  since 
we  are  assuming  that  /  is  of  class  Ck+1,  we  have  that  f(k+x\t)  is  continuous  and 
therefore,  bounded  for  t  between  a  and  x  (i.e.,  that  \f{k+l\t)\  <  M  for  some 
constant  M).  Thus, 


\Rk(x,a)\  < 


r  (x  -  tf 

Ja  k\ 


f(k+\t)dt 


<  ± 


f 


(x  -  tf 


k\ 


fik+l)(t) 


dt, 


where  the  plus  sign  applies  if  x  >  a  and  the  negative  sign  if  x  <  a, 


Cx  M 


x  -  a\k  dt 


M 


\k+\ 


Thus, 


Rk(x,  a) 


(x  —  a)k 


M 


as  x  —>  a,  as  desired. 


260       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


Proof  of  Proposition  1.2  We  establish  Proposition  1.2  by  means  of  a  general 
version  of  the  mean  value  theorem  for  integrals.  This  theorem  states  that  for 
continuous  functions  g  and  h  such  that  h  does  not  change  sign  on  [a,  b]  (i.e., 
either  h(t)  >  0  on  [a,  b]  or  h(t)  <  0  on  [a,  b]),  there  is  some  number  z  between 
a  and  b  such  that 

f  g(t)h(t)dt  =  g(z)  f  h(t)dt. 

(We  omit  the  proof  but  remark  that  this  theorem  is  a  consequence  of  the  interme- 
diate value  theorem.)  Applying  this  result  to  formula  (14)  with  g(t)  =  f(k+^(t) 
and  h(t)  =  (x  —  t)k /k\,  we  find  that  there  must  exist  some  z  between  a  and  x 
such  that 


i)  =  f(k+\z)\X{^-^dt=f^\z) 


k\ 


(x  -  t) 


(k+iy. 


(k  +  iy. 


(x  —  a) 


k+l 


Proof  of  Theorem  1.5  As  in  the  proof  of  Theorem  1 . 1 ,  we  establish  Theorem  1 .5 
under  the  stronger  assumption  that  /  is  of  class  C3 .  Begin  by  setting  h  =  x  —  a, 
so  that  x  =  a  +  h,  and  consider  a  and  h  to  be  fixed.  We  define  the  one-variable 
function  F  by  F(t)  =  /(a  +  th).  Since  /  is  assumed  to  be  of  class  C3  on  an 
open  set  X,  if  we  take  x  sufficiently  close  to  a,  then  F  is  of  class  C3  on  an  open 
interval  containing  [0,  1].  Thus,  Theorem  1.1  with  k  =  2,  a  =  0,  and  x  =  1  may 
be  applied  to  give 


F(l)  =  F(0)  +  F'(0)(1  -  0)  +  ^9(1  -  0)2  +  R2(l,  0) 
F"(0) 

=  F(0)  +  F'(0)  +  — ^  +  R2(l,  0), 


(15) 


where R2(l,  0)  =  /J  ^-F"'(0^.Now  we  use  the  chain  rule  to  calculate  deriva- 
tives of  F  in  terms  of  partial  derivatives  of  /: 


.7  =  1 


hi  =  ^  fXiXj(a  +  th)hihj; 


F'(t)  =  D/(a  +  th)h  =  J2  U ■(»  +  fh)^ 
F"(r)  =  £ 
F"'(t)  = 

k=\ 

Thus,  (15)  becomes 

n  j  n 

/(a  +  h)  =  /(a)  +      fx,Wi  +  ~  E  UM)hihJ 


J2  /u,,,;(a  •  /h)/;,/;, 
.W=i 


fyfc  =  X!  fxiXjxM  +  t^hihjhk 
i,j,k=\ 


+ 


n  1 


(l  -  o2 


fXiXjXk(a  +  th)hihjhkdt, 


4.1  |  Differentials  and  Taylor's  Theorem  261 


or,  equivalently, 

n  \  n 

/(x)  =  /(a)  +      /*,(»)(*;  "  ad  +~Y1  fxitjW&i  ~  ai)(xi  ~  ai) 

+  #2(x,  a), 
where  the  multivariable  remainder  is 

R2(x,a)=  V   /    y——LfXiX]Xl(a  +  th)hihjhkdt. 

>,j,k=\ Jo  l 


(16) 


We  must  still  show  that  |i?2(x,  a)|/||x  —  a||2  ->  0  as  x  ->  a,  or,  equivalently, 
that  |^?2(x,  a)|/|h||2  ->  0  as  h  ->  0.  To  demonstrate  this,  note  that,  for  a  and  h 
fixed  the  expression  (1  —  t)2  fXjXjXk(st  +  fh)  is  continuous  for  f  in  [0,  1]  (since  / 
is  assumed  to  be  of  class  C3),  hence  bounded.  In  addition,  for  i  =  1, . . . ,  n,  we 


have  that  |/z,-|  <  ||h||. 

Hence, 

|tf2(x,a)|  = 

< 

;i  1 

< 

n  «1 

(i  -  o2 

2 

(1  "  tf 


fxtXjxM  +  th)hihjhk  dt 
fXiXjXt(a  +  th)hjhjhk  dt 
M||/i||3  dt  =  n3M||h||3  =  n3M||x  -  a| 


Thus, 


|i?2(x,  a)|  , 

v     '1  <«3M||x-a| 


x-  a 


0 


as  x  ->  a. 

Finally,  we  remark  that  entirely  similar  arguments  may  be  given  to  establish 
results  for  Taylor  polynomials  of  orders  higher  than  two.  ■ 

Lagrange's  formula  for  the  remainder  (see  page  257)  Using  the  function 
F(t)  =  /(a  +  fh)  defined  in  the  proof  of  Theorem  1.5,  Proposition  1.2  implies 
that  there  must  be  some  number  c  between  0  and  1  such  that  the  one-variable 
remainder  is 

F"'(c)  - 
tf2(l,0)=^-Al-0)3. 

Now,  the  remainder  term  Rz(I,  0)  from  Proposition  1.2  is  precisely  ^(x,  a)  in 
Theorem  1.5  and 

n  n 

F"'(c)=         fXiXiXt{*  +  c\v)hihjhk=  fXiXjXk(z)hihjhk, 

i,j,k=\  i,j,k=l 

where  z  =  a  +  ch.  Since  c  is  between  0  and  1 ,  the  point  z  lies  on  the  line  segment 
joining  a  and  x  =  a  +  h,  and  so 

1 


^2(x,  a)=—  fXiXjXk(z)hihjhk 


i,j,k=l 


which  is  the  result  we  desire.  The  derivation  of  the  formula  for  Rk(x,  a)  for  k  >  2 
is  analogous.  ■ 


262       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 

4.1  Exercises 


In  Exercises  1—7,  find  the  Taylor  polynomials  of  given  order 
k  at  the  indicated  point  a. 

1.  f(x)  =  e2x,a  =  0,k  =  A 

2.  fix)  =  ln(l  +  x),  a  =  0,k  =  3 

3.  f(x)=  l/x2,a  =  l,k  =  4 

4.  f(x)  =  Jx,a  =  l,k  =  3 

5.  f(x)  =  </x,  a  =  9,  k  =  3 

6.  f(x)  =  sinx,  a  =  0,  k  =  5 

7.  f(x)  =  sinx,  a  =  n/2,  k  =  5 

In  Exercises  8— 15,  find  the  first-  and  second-order  Taylor  poly- 
nomials for  the  given  function  f  at  the  given  poin  t  a. 

8.  /(x,y)  =  l/(x2  +  y2  +  l),a  =  (0,0) 

9.  f(x,  y)  =  l/(x2  +  y2  +  1),  a  =  (1,  -1) 

10.  f(x,y)  =  e2x+y,a  =  (0,  0) 

11.  f(x,  y)  =  e2x  cos3y,  a  =  (0,  jr) 

12.  /(x,  y,  z)  =  ye3x  +  ze2-\  a  =  (0,  0,  2) 

13.  /(jc,  y,  z)  =  xy-  3y2  +  2xz,  a  =  (2,  -1,  1) 

14.  /(x,     z)  =  l/(x2  +  y2  +  z2  +  1),  a  =  (0,  0,  0) 

15.  f(x,  y,  z)  =  sinxyz,  a  =  (0,  0,  0) 

In  Exercises  16-20,  calculate  the  Hessian  matrix  ///(a)  for 
the  indicated  function  f  at  the  indicated  point  a. 

16.  f(x,y)=  l/(x2  +  y2+l),a  =  (0,0) 

17.  /(x,  y)  =  cosx  siny,  a  =  (tt/4,  it/3) 

18.  f(x,y,z)=  -£=,a  =  (l,2,-4) 

19.  /(x,  y,  z)  =  x3  +  x2y  -  yz2  +  2z\  a  =  (1,  0,  1) 

20.  f(x,  y,  z)  =  e2x~^  sin5z,  a  =  (0,  0,  0) 

21 .  For  /  and  a  as  given  in  Exercise  8,  express  the  second- 
order  Taylor  polynomial  pi(x,  y),  using  the  derivative 
matrix  and  the  Hessian  matrix  as  in  formula  (10)  of 
this  section. 

22.  For  /  and  a  as  given  in  Exercise  1 1 ,  express  the  second- 
order  Taylor  polynomial  pi{x,  y),  using  the  derivative 
matrix  and  the  Hessian  matrix  as  in  formula  (10)  of 
this  section. 

23.  For  /  and  a  as  given  in  Exercise  12,  express  the  second- 
order  Taylor  polynomial  p2(x,  y,  z),  using  the  deriva- 
tive matrix  and  the  Hessian  matrix  as  in  formula  (10) 
of  this  section. 


24.  For  /  and  a  as  given  in  Exercise  1 9,  express  the  second- 
order  Taylor  polynomial  pi(x,  y,  z),  using  the  deriva- 
tive matrix  and  the  Hessian  matrix  as  in  formula  (10) 
of  this  section. 

25.  Consider  the  function 

f(xux2,...,x,,)  =  ex>+2x>+-+"x». 

(a)  Calculate  D/(0,  0, . . . ,  0)  and  Hf(0,  0, . . . ,  0). 

(b)  Determine  the  first-  and  second-order  Taylor  poly- 
nomials of  /  at  0. 

(c)  Use  formulas  (3)  and  ( 10)  to  write  the  Taylor  poly- 
nomials in  terms  of  the  derivative  and  Hessian 
matrices. 

26.  Find  the  third-order  Taylor  polynomial  pi(x,  y,  z)  of 

f(x,y,z)  =  ex+2y+3z 

at  (0,  0,  0). 

27.  Find  the  third-order  Taylor  polynomial  of 

f{x,  y,  z)  =  x4  +  x3y  +  2y3  -  xz2  +  x2y  +  3xy  -  z  +  2 

(a)  at  (0,  0,  0). 

(b)  at  (1,-1,0). 

Determine  the  total  differential  of  the  functions  given  in 
Exercises  28—32. 

28.  /(x,y)  =  x2y3 

29.  f{x,  y,  z)  =  x2  +  3y2  -  2z3 

30.  fix,  y,  z)  =  cos  (xyz) 

31.  fix,  y,  z)  =  ex  cosy  +  ey  sin  z 

32.  fix,  y,  z)  =  l/^Jxyz 

33.  Use  the  fact  that  the  total  differential  df  approximates 
the  incremental  change  A/  to  provide  estimates  of  the 
following  quantities: 

(a)  (7.07)2(1.98)3 

(b)  l/y(4.1)(1.96)(2.05) 

(c)  (1. 1)  cos  ((jr  -0.03X0.12)) 

34.  Near  the  point  (1,  —2,  1),  is  the  function  g(x,  y,  z)  = 
x3  —  2xy  +  x2z  +  7z  most  sensitive  to  changes  in  x, 
y,  or  z? 

35.  To  which  entry  in  the  matrix  is  the  value  of  the 
determinant 

2  3 
-1  5 

most  sensitive? 


4.2  |  Extrema  of  Functions  263 


36.  If  you  measure  the  radius  of  a  cylinder  to  be  2  in,  with 
a  possible  error  of  ±0. 1  in,  and  the  height  to  be  3  in, 
with  a  possible  error  of  ±0.05  in,  use  differentials  to 
determine  the  approximate  error  in 

(a)  the  calculated  volume  of  the  cylinder. 

(b)  the  calculated  surface  area. 

37.  A  can  of  mushrooms  is  currently  manufactured  to  have 
a  diameter  of  5  cm  and  a  height  of  12  cm.  The  man- 
ufacturer plans  to  reduce  the  diameter  by  0.5  cm.  Use 
differentials  to  estimate  how  much  the  height  of  the 
can  would  need  to  be  increased  in  order  to  keep  the 
volume  of  the  can  the  same. 

38.  Consider  a  triangle  with  sides  of  lengths  a  and  b  that 
make  an  interior  angle  0 . 

(a)  If  a  =  3,  b  =  4,  and  0  =  tt/3,  to  changes  in  which 
of  these  measurements  is  the  area  of  the  triangle 
most  sensitive? 

(b)  If  the  length  measurements  in  part  (a)  are  in  error 
by  as  much  as  5%  and  the  angle  measurement  is 
in  error  by  as  much  as  2%,  estimate  the  resulting 
maximum  percentage  error  in  calculated  area. 

39.  To  estimate  the  volume  of  a  cone  of  radius  approx- 
imately 2  m  and  height  approximately  6  m,  how  ac- 
curately should  the  radius  and  height  be  measured  so 
that  the  error  in  the  calculated  volume  estimate  does 


not  exceed  0.2  m3?  Assume  that  the  possible  errors  in 
measuring  the  radius  and  height  are  the  same. 

40.  Suppose  that  you  measure  the  dimensions  of  a  block 
of  tofu  to  be  (approximately)  3  in  by  4  in  by  2  in. 
Assuming  that  the  possible  errors  in  each  of  your  mea- 
surements are  the  same,  about  how  accurate  must  your 
measurements  be  so  that  the  error  in  the  calculated 
volume  of  the  tofu  is  not  more  than  0.2  in3?  What  per- 
centage error  in  volume  does  this  represent? 

41 .  (a)  Calculate  the  second-order  Taylor  polynomial  for 

f(x,  y)  =  cosx  siny  at  the  point  (0,  ?r/2). 

(b)  If  h  =  (huh2)  =  (x,  y)  -  (0,  tt/2)  is  such  that 
\hi  \  and  \h2\  are  no  more  than  0.3,  estimate  how 
accurate  your  Taylor  approximation  is. 

42.  (a)  Determine  the  second-order  Taylor  polynomial  of 

f(x,  y)  =  ex+2y  at  the  origin. 

(b)  Estimate  the  accuracy  of  the  approximation  if  \x\ 
and  \  y\  are  no  more  than  0.1. 

43.  (a)  Determine  the  second-order  Taylor  polynomial  of 

f(x,  y)  =  e2x  cos  y  at  the  point  (0,  rr/2). 

(b)  If  h  =  (huh2)  =  (x,  y)  -  (0,  tt/2)  is  such  that 
\h\\  <  0.2  and  \h2\  <  0.1,  estimate  the  accuracy 
of  the  approximation  to  /  given  by  your  Taylor 
polynomial  in  part  (a). 


4.2   Extrema  of  Functions 

The  power  of  calculus  resides  at  least  in  part  in  its  role  in  helping  to  solve  a  wide 
variety  of  optimization  problems.  With  any  quantity  that  changes,  it  is  natural  to 
ask  when,  if  ever,  does  that  quantity  reach  its  largest,  its  smallest,  its  fastest  or 
slowest?  You  have  already  learned  how  to  find  maxima  and  minima  of  a  function 
of  a  single  variable,  and  no  doubt  you  have  applied  your  techniques  to  a  number  of 
situations.  However,  many  phenomena  are  not  appropriately  modeled  by  functions 
of  only  one  variable.  Thus,  there  is  a  genuine  need  to  adapt  and  extend  optimization 
methods  to  the  case  of  functions  of  more  than  one  variable.  We  develop  the 
necessary  theory  in  this  section  and  the  next  and  explore  a  few  applications  in  §4.4. 

Critical  Points  of  Functions   

Let  X  be  open  in  R"  and  /:XcR"^Ra  scalar- valued  function. 


DEFINITION  2.1  We  say  that  /  has  a  local  minimum  at  the  point  a  in 
X  if  there  is  some  neighborhood  U  of  a  such  that  f(x)  >  /(a)  for  all  x 
in  U.  Similarly,  we  say  that  /  has  a  local  maximum  at  a  if  there  is  some 
neighborhood  U  of  a  such  that  /(x)  <  /(a)  for  all  x  in  U. 


Figure  4.1 2  The  graph  of  When  n  =  2,  local  extrema  of  f(x,  y)  are  precisely  the  pits  and  peaks  of  the 

z  =  f(x,  y).  surface  given  by  the  graph  of  z  =  f(x,  y),  as  suggested  by  Figure  4.12. 


Max. 


264       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


We  emphasize  our  use  of  the  adjective  "local."  When  a  local  maximum  of 
a  function  /  occurs  at  a  point  a,  this  means  that  the  values  of  /  at  points  near 
a  can  be  no  larger,  not  that  all  values  of  /  are  no  larger.  Indeed,  /  may  have 
local  maxima  and  no  global  (or  absolute)  maximum.  Consider  the  graphs  in 
Figure  4.13.  (Of  course,  analogous  comments  apply  to  local  and  global  minima.) 


Figure  4.1 3  Examples  of  local  and  global  maxima. 


Recall  that,  if  a  differentiable  function  of  one  variable  has  a  local  extremum 
at  a  point,  then  the  derivative  vanishes  there  (i.e.,  the  tangent  line  to  the  graph 
of  the  function  is  horizontal).  Figures  4.12  and  4.13  suggest  strongly  that,  if  a 
function  of  two  variables  has  a  local  maximum  or  minimum  at  a  point  in  the 
domain,  then  the  tangent  plane  at  the  corresponding  point  of  the  graph  must  be 
horizontal.  Such  is  indeed  the  case,  as  the  following  general  result  (plus  formula 
(4)  of  §2.3)  implies. 


THEOREM  2.2  Let  X  be  open  in  R"  and  let  /:XcR"^R  be  differentiable. 
If  /  has  a  local  extremum  at  a  6  X,  then  Df(a)  =  0. 


PROOF  Suppose,  for  argument's  sake,  that  /  has  a  local  maximum  at  a.  Then  the 
one-variable  function  F  defined  by  F(t)  =  /(a  + 1  h)  must  have  a  local  maximum 
at  t  =  0  for  any  h.  (Geometrically,  the  function  F  is  just  the  restriction  of  /  to  the 
line  through  a  parallel  to  h  as  shown  in  Figure  4. 14.)  From  one-variable  calculus, 
we  must  therefore  have  F'(0)  =  0.  By  the  chain  rule 


F'(t)  =  -[/(a  +  th)]  =  D/(a  +  th)h  =  V/(a  +  th)  •  h. 
at 


1  \ 
1  k 

1  /  \ 

1  1 

1  /  1 

1 

 1  J 

Graph  of  / 
restricted  to  line 


Figure  4.1 4  The  graph  of  /  restricted  to  a  line. 


4.2  |  Extrema  of  Functions  265 


Hence, 


/-o 

X/<o 

/  \ 

/  \ 
1  \ 

f-o 

v/ 
/\ 
/  \ 

/f>0\ 

\  f>0/ 

\  1 
\  / 
\  / 

\  / 
\  / 

\  / 

Figure  4.1 5  The  function  /  is 
strictly  positive  on  the  shaded 
region,  strictly  negative  on  the 
unshaded  region,  and  zero  along 
the  lines  y  =  ±x. 


0  =  F'(0)  =  D/(a)h  =  fx^)hx  +  /a(a)A2  +  •  •  •  +  A„(a)fc„. 

Since  this  last  result  must  hold  for  all  h  e  R",  we  find  that  by  setting  h  in  turn 
equal  to  (1,  0,  . . . ,  0),  (0,  1,  0, . . . ,  0),  . . . ,  (0, . . . ,  0,  1),  we  have 

fXl{*)=         =  -••  =  /*„  (a)  =  0. 

Therefore,  Df(a)  =  0,  as  desired.  ■ 

A  point  a  in  the  domain  of  /  where  Df(a)  is  either  zero  or  undefined  is  called 
a  critical  point  of  /.  Theorem  2.2  says  that  any  extremum  of  /  must  occur  at  a 
critical  point.  However,  it  is  by  no  means  the  case  that  every  critical  point  must 
be  the  site  of  an  extremum. 

EXAMPLE  1  If  f(x,  y)  =  x2  -  y2,  then  Df(x,  y)  =  [  2x  -2y  ]  so  that, 
clearly,  (0,  0)  is  the  only  critical  point.  However,  neither  a  maximum  nor  a  mini- 
mum occurs  at  (0,  0).  Indeed,  inside  every  open  disk  centered  at  (0,  0),  no  matter 
how  small,  there  are  points  for  which  f(x,  y)  >  f(0,  0)  =  0  and  also  points  where 
f(x,  y)  <  /(O,  0).  (See  Figure  4.15.)  ♦ 

This  type  of  critical  point  is  called  a  saddle  point.  Its  name  derives  from  the 
fact  that  the  graph  of  z  =  f(x ,  y)  looks  somewhat  like  a  saddle.  (See  Figure  4.16.) 


Figure  4.1 6  A  saddle  point. 


EXAMPLE  2    Let  f(x,  y)  =  ^x1  +  y2.  The  domain  off  is  all  of  R2.  We  com- 

2x  2y 


pute  that  Df(x,  y) 


;  note  that  Df  is  unde- 


3(x2  +  y2)2/3       3(x2  +  _y2)2/3  _ 

fined  at  (0,  0)  and  nonzero  at  all  other  (x,  y)  e  R2.  Hence,  (0,  0)  is  the  only 
critical  point.  Since  f(x,  y)  >  0  for  all  (x,  y)  and  has  value  0  only  at  (0,  0),  we 
see  that  /  has  a  unique  (global)  minimum  at  (0,  0).  ♦ 


The  Nature  of  a  Critical  Point:  The  Hessian  Criterion 

We  illustrate  our  current  understanding  regarding  extrema  with  the  following 
example: 


EXAMPLE  3   We  find  the  extrema  of 

f(x,  y)  =  x2  +  xy  +  y2  +  2x-  2y  +  5. 


266       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


/(«) 


Figure  4.17  An 

upward-opening  parabola. 


Figure  4.18  A 

downward-opening  parabola. 


Since  /  is  a  polynomial,  it  is  differentiable  everywhere,  and  Theorem  2.2  implies 
that  any  extremum  must  occur  where  df/dx  and  df/dy  vanish  simultaneously. 
Thus,  we  solve 


Bf 
dx 
Bf 
dy 


=  2x  +  y  +  2  =  0 
=  x+2y-2  =  0 


and  find  that  the  only  solution  is  x  =  —2,  y  =  2.  Consequently,  (—2,  2)  is  the 
only  critical  point  of  this  function. 

To  determine  whether  (—2,  2)  is  a  maximum  or  minimum  (or  neither),  we 
could  try  graphing  the  function  and  drawing  what  we  hope  would  be  an  obvious 
conclusion.  Of  course,  such  a  technique  does  not  extend  to  functions  of  more  than 
two  variables,  so  a  graphical  method  is  of  limited  value  at  best.  Instead  we'll  see 
how  /  changes  as  we  move  away  from  the  critical  point: 

A/  =  /(-2  +  A,2  +  *)-/(-2,2) 

=  [(-2  +  hf  +  (-2  +  h)(2  +  k)  +  {2  +  kf 

+  2(-2  +  /!)-2(2  +  £)  +  5]-  1 
=  h2  +  hk  +  k2. 

If  the  quantity  Af  =  h2  +  hk  +  k2  is  nonnegative  for  all  small  values  of  h  and  k, 
then  (—2,  2)  yields  a  local  minimum.  Similarly,  if  Af  is  always  nonpositive,  then 
(—2,  2)  must  yield  a  local  maximum.  Finally,  if  Af  is  positive  for  some  values 
of  h  and  k  and  negative  for  others,  then  (—2,  2)  is  a  saddle  point.  To  determine 
which  possibility  holds,  we  complete  the  square: 


Af 


h2  +  hk  +  k2 


h2  +  hk  + 


,k  +  .  k 


(h  +  \kf  +  Ik2. 


Thus,  Af  >  0  for  all  values  of  h  and  k,  so  (—2,  2)  necessarily  yields  a  local 
minimum.  ♦ 

Example  3  with  its  attendant  algebra  clearly  demonstrates  the  need  for  a  better 
way  of  determining  when  a  critical  point  yields  a  local  maximum  or  minimum  (or 
neither).  In  the  case  of  a  twice  differentiable  function /:XcR^R,  you  already 
know  a  quick  method  namely,  consideration  of  the  sign  of  the  second  derivative. 
This  method  derives  from  looking  at  the  second-order  Taylor  polynomial  of  / 
near  the  critical  point  a,  namely, 

f"(a)  , 

f(x)  «  p2(x)  =  f(a)  +  /  (a)(x  -a)-\  —  (x  -  a) 

=  /(a)H  ^—(x  ~  a)  , 

since  /'  is  zero  at  the  critical  point  a  of  /.  If  f"{a)  >  0,  the  graph  of  y  =  pi(x) 
is  an  upward-opening  parabola,  as  in  Figure  4.17,  whereas  if  f"{a)  <  0,  then 
the  graph  of  y  =  pi(x)  looks  like  the  one  shown  in  Figure  4.18.  If  f"(a)  =  0, 
then  the  graph  of  y  =  pi(x)  is  just  a  horizontal  line,  and  we  would  need  to  use 
a  higher-order  Taylor  polynomial  to  determine  if  /  has  an  extremum  at  a.  (You 
may  recall  that  when  f"(a)  =  0,  the  second  derivative  test  from  single-variable 
calculus  gives  no  information  about  the  nature  of  the  critical  point  a.) 
The  concept  is  similar  in  the  context  of  n  variables.  Suppose  that 

/(*)=  f(xi,x2,  . . .  ,xn) 


4.2  |  Extrema  of  Functions  267 


is  of  class  C2  and  that  a  =  (a\,  a2, . . . ,  a„)  is  a  critical  point  of  /.  Then  the 
second-order  Taylor  approximation  to  /  gives 

a/  =  /(x)  -  m  «  P2(x)  -  m 

=  Z)/(a)(x  -  a)  +  i(x  -  a)r///(a)(x  -  a) 

when  x  a.  (See  Theorem  1.5  and  formula  (10)  in  §4.1.)  Since  /  is  of  class  C2 
and  a  is  a  critical  point,  all  the  partial  derivatives  vanish  at  a,  so  that  we  have 
D/(a)  =  0  and  hence, 


A/«i(x-a)r///(a)(x-a). 


(1) 


The  approximation  in  (1)  suggests  that  we  may  be  able  to  see  whether  the  in- 
crement A /  remains  positive  (respectively,  remains  negative)  for  x  near  a  and 
hence,  whether  /  has  a  local  minimum  (respectively,  a  local  maximum)  at  a  by 
seeing  what  happens  to  the  right  side. 

Note  that  the  right  side  of  (1),  when  expanded,  is  quadratic  in  the  terms 
(xj  —  cii).  More  generally,  a  quadratic  form  in  h\,  h2,  ■ . . ,  h„  is  a  function  Q 
that  can  be  written  as 

n 

Q(hu  h2, hn)  =  ^  bijhihj, 

where  the  bij 's  are  constants.  The  quadratic  form  Q  can  also  be  written  in  terms 
of  matrices  as 


Q(h)=[hl  h2 


~  bu 

b\2  ■ 

■  bl„ 

~  hx  ~ 

h„] 

b2\ 

b22  ■ 

■  b2n 

h2 

_  K\ 

bnl  ■ 

■  h 

unn  - 

_  K  _ 

hrfih, 


(2) 


where  B  =  (pij).  Note  that  the  function  Q  is  unchanged  if  we  replace  all  bjj  with 
^{bij  +  bjj).  Hence,  we  may  always  assume  that  the  matrix  B  associated  to  Q  is 
symmetric,  that  is,  that  bjj  =  bj,  (or,  equivalently,  that  BT  =  B).  Ignoring  the 
factor  of  1  /2,  we  see  that  the  right  side  of  (1)  is  the  quadratic  form  in  h  =  x  —  a, 
corresponding  to  the  matrix  B  =  Hf(a). 

A  quadratic  form  Q  (respectively,  its  associated  symmetric  matrix  B)  is  said 
to  be  positive  definite  if  Q(h)  >  0  for  all  h  ^  0  and  negative  definite  if  Q(h)  <  0 
for  all  h  ^  0.  Note  that  if  Q  is  positive  definite,  then  Q  has  a  global  minimum  (of 
0)  at  h  =  0.  Similarly,  if  Q  is  negative  definite,  then  Q  has  a  global  maximum  at 
h  =  0. 

The  importance  of  quadratic  forms  to  us  is  that  we  can  judge  whether  /  has 
a  local  extremum  at  a  critical  point  a  by  seeing  if  the  quadratic  form  in  the  right 
side  of  ( 1)  has  a  maximum  or  minimum  at  x  =  a.  The  precise  result,  whose  proof 
is  given  in  the  addendum  to  this  section,  is  the  following: 


THEOREM  2.3  Let  U  c  R"  be  open  and  /:  U  R  a  function  of  class  C2. 
Suppose  that  a  £  U  is  a  critical  point  of  /. 

1.  If  the  Hessian  Hf(a)  is  positive  definite,  then  /  has  a  local  minimum  at  a. 

2.  If  the  Hessian  Hf(a)  is  negative  definite,  then  /  has  a  local  maximum  at  a. 

3.  If  det  Hf(a)     0  but  Hf(a)  is  neither  positive  nor  negative  definite,  then  / 
has  a  saddle  point  at  a. 


268       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


In  view  of  Theorem  2.3,  the  issue  thus  becomes  to  determine  when  the  Hessian 
Hf(a)  is  positive  or  negative  definite.  Fortunately,  linear  algebra  provides  an 
effective  means  for  making  such  a  determination,  which  we  state  without  proof. 
Given  a  symmetric  matrix  B  (which,  as  we  have  seen,  corresponds  to  a  quadratic 
form  Q),  let  B^,  for  k  =  1, . . . ,  n,  denote  the  upper  leftmost  k  x  k  submatrix  of 
B.  Calculate  the  following  sequence  of  determinants: 

bn  b22 


det  5, 


det  B7 


det  B3  = 


bn  bi2  bn 
bi\  bi2  bn 
b-i\   bi2  b3i 


det£„  =  det  5. 


If  this  sequence  consists  entirely  of  positive  numbers,  then  B  and  Q  are  positive 
definite.  If  this  sequence  is  such  that  det  B^  <  0  for  k  odd  and  det  B^  >  0  for  k 
even,  then  B  and  Q  are  negative  definite.  Finally,  if  det  B  ^  0,  but  the  sequence 
of  determinants  det  B\ ,  det  B2, . . . ,  det  B„  is  neither  of  the  first  two  types,  then  B 
and  Q  are  neither  positive  nor  negative  definite.  Combining  these  remarks  with 
Theorem  2.3,  we  can  establish  the  following  test  for  local  extrema: 


Second  derivative  test  for  local  extrema.  Given  a  critical  point  a  of  a  func- 
tion /  of  class  C2,  look  at  the  Hessian  matrix  evaluated  at  a: 


tf/(a) 


/*,*,(»)  fxlX2(*)  ■■■  /*,*„(») 
fX2xM)  fx2x2(a)  •••  /x2x„(a) 


From  the  Hessian,  calculate  the  sequence  of  principal  minors  of  Hf(a). 
This  is  the  sequence  of  the  determinants  of  the  upper  leftmost  square  sub- 
matrices  of  Hf(a).  More  explicitly,  this  is  the  sequence  d\,  d2, . . . ,  dn, 
where  dt  =  det  Ht,  and  Hk  is  the  upper  leftmost  k  x  k  submatrix  of  Hf(a). 
That  is, 

d\ 
di 


/x,x,(a), 

/««(») 

/v!*2(a) 

fx2xM) 

fx2x2(&) 

1 

fxixM) 

fxlxM 

fxlxM 

fx2xM) 

fx2xM 

fx2xM 

/v3*,(a) 

/v3*2(a) 

inu(a) 

di=    fx2xM)     fx2x2(a)     fx2x3(a)   ,  ...,d„  =  \Hf(a)\. 


The  numerical  test  is  as  follows: 
Assume  that  d„  =  det  Hf(a)  ^  0. 

1.  If  dk  >  0  for  k  =  1 , 2, . . . ,  n,  then  /  has  a  local  minimum  at  a. 

2.  If  dk  <  0  for  k  odd  and  dk  >  0  for  k  even,  then  /  has  a  local  maximum 
at  a. 

3.  If  neither  case  1  nor  case  2  holds,  then  /  has  a  saddle  point  at  a. 

In  the  event  that  det  ///(a)  =  0,  we  say  that  the  critical  point  a  is  degenerate 
and  must  use  another  method  to  determine  whether  or  not  it  is  the  site  of  an 
extremum  of  /. 


4.2  |  Extrema  of  Functions  269 


EXAMPLE  4   Consider  the  function 


fxx 

fxy 

2  1 

.  fyx 

fyy  _ 

1  2 

minors 

is  d 

i  =  /«(-2,  2) 

/(*,  y)  =  x2  +  xy  +  y2  +  2x  -  2y  +  5 

in  Example  3.  We  have  already  seen  that  (—2,  2)  is  the  only  critical  point.  The 
Hessian  is 


Hf(x,  y)  = 


\Hf(—2,  2)|  =  3  (>  0).  Hence,  /  has  a  minimum  at  (—2,  2),  as  we  saw  before, 
but  this  method  uses  less  algebra.  ♦ 

EXAMPLE  5  (Second  derivative  test  for  functions  of  two  variables)  Let  us 

generalize  Example  4.  Suppose  that  f(x,  y)  is  a  function  of  two  variables  of  class 
C2  and  further  suppose  that  /  has  a  critical  point  at  a  =  (a,  b).  The  Hessian 
matrix  of  /  evaluated  at  {a,  b)  is 


Hf(a,b) 


fxx(a,b)  fxy(a,b) 
fxy(a,b)  fyy(a,b) 


Note  that  we  have  used  the  fact  that  fxy  =  fyx  (since  /  is  of  class  C2)  in  con- 
structing the  Hessian.  The  sequence  of  principal  minors  thus  consists  of  two 
numbers: 

d\  =  fxx(a,b)    and    d2  =  fxx(a,  b)fyy(a,  b)  -  fxy(a,  bf . 
Hence,  in  this  case,  the  second  derivative  test  tells  us  that 

1.  /  has  a  local  minimum  at  (a,  b)  if 

fxx(a,  b)  >  0    and    fxx(a,  b)fyy(a,  b)  -  fxy(a,  bf  >  0. 

2.  /  has  a  local  maximum  at  {a,  b)  if 

fxx(a,  b)  <  0    and    fxx(a,  b)fyy(a,  b)  -  fxy(a,  bf  >  0. 

3.  /  has  a  saddle  point  at  (a,  b)  if 

fxxia,  b)fyy(a,  b)  -  fxy(a,  bf  <  0. 

Note  that  if  fxx(a,  b)fyy(a,  b)  —  fxy(a,  bf  =  0,  then  /  has  a  degenerate  critical 
point  at  (a,  b)  and  we  cannot  immediately  determine  if  {a,  b)  is  the  site  of  a  local 
extremum  of  /.  ♦ 

EXAMPLE  6  Let  f(x,  y,  z)  =  x3  +  xy2  +  x2  +  y2  +  3z2.  To  find  any  local 
extrema  of  /,  we  must  first  identify  the  critical  points.  Thus,  we  solve 

Df(x,y,z)=  [3x2  +  y2  +  2x   2xy  +  2y   6z]  =  [0  0  0]  . 

From  this,  it  is  not  hard  to  see  that  there  are  two  critical  points:  (0,  0,  0)  and 
(-§,  0,0).  The  Hessian  of/  is 


Hf(x,y,z)  = 


6x  +  2 

2y 
o 


2y 
2x  +  2 
0 


270       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


At  the  critical  point  (0,  0,  0),  we  have 


Hf(0,  0,  0)  = 


2  0  0 
0  2  0 
0     0  6 


and  its  sequence  of  principal  minors  is  d\  =2,  d2  =  4,  dj  =  24.  Since  these 
determinants  are  all  positive,  we  conclude  that  /  has  a  local  minimum  at  (0,0,0). 
At  (-§,0,0),  we  calculate  that 


ff/(--,0,0 


-2  0  0 
0  |  0 
0     0  6 


The  sequence  of  minors  is  —2,  —  |,  — 8.  Hence,  /  has  a  saddle  point  at  (— |,  0,  0). 


+ 

h>0 

+ 

+ 

+ 

 1 — 

-1 

0 

+ 

 1  

1 

+ 

+ 

+ 

Figure  4.1 9  Away  from  the 
origin,  the  function  h  of  Example  7 
is  negative  along  the  x-axis  and 
positive  along  the  _v-axis. 


EXAMPLE  7  To  get  a  feeling  for  what  happens  in  the  case  of  a  degenerate 
critical  point  (i.e.,  a  critical  point  a  such  that  det  Hf(a)  =  0),  consider  the  three 
functions 


/(x,y)  =  x4  +  x2  +  y4, 


g(x,  y)  =  -x4  -  x2 


and 


h(x,  y)  =  x 


x2  +  y4. 


We  leave  it  to  you  to  check  that  the  origin  (0,  0)  is  a  degenerate  critical  point 
of  each  of  these  functions.  (In  fact,  the  Hessians  themselves  look  very  similar.) 
Since  /  is  0  at  (0,  0)  and  strictly  positive  at  all  (x,  y)  ^  (0,  0),  we  see  that  / 
has  a  strict  minimum  at  the  origin.  Similar  reasoning  shows  that  g  has  a  strict 
maximum  at  the  origin.  For  h,  the  situation  is  slightly  more  complicated.  Along 
the  y-axis,  we  have  h(0,  y)  =  y4,  which  is  zero  at  y  =  0  (the  origin)  and  strictly 
positive  everywhere  else.  Along  the  x-axis, 


h(x,0)  =  x4 


x2  =  x2(x 


1)(JC  +  1). 


For  —  1  <  x  <  1  and  x  ^  0,  h(x,  0)  <  0.  We  have  the  situation  depicted  in  Fig- 
ure 4. 19.  Thus,  every  neighborhood  of  (0,  0)  contains  some  points  (x,  y)  where  h 
is  positive  and  also  some  points  where  h  is  negative.  Therefore,  h  has  a  saddle  point 
at  the  origin.  The  "moral  of  the  story"  is  that  a  degenerate  critical  point  can  exhibit 
any  type  of  behavior,  and  more  detailed  consideration  of  the  function  itself,  rather 
than  its  Hessian,  is  necessary  to  understand  its  nature  as  a  site  of  an  extremum.  ♦ 


Global  Extrema  on  Compact  Regions 

Thus  far  our  discussion  has  been  limited  to  consideration  of  only  local  extrema. 
We  have  said  nothing  about  how  to  identify  global  extrema,  because  there  really  is 
no  general,  effective  method  for  looking  at  an  arbitrary  function  and  determining 
whether  and  where  it  reaches  an  absolute  maximum  or  minimum  value.  For  the 
purpose  of  applications,  where  finding  an  absolute  maximum  or  minimum  is 
essential,  such  a  state  of  affairs  is  indeed  unfortunate.  Nonetheless,  we  can  say 
something  about  global  extrema  for  functions  defined  on  a  certain  type  of  domain. 


4.2  |  Extrema  of  Functions  271 


Figure  4.20  Compact  regions. 


DEFINITION  2.4  A  subset  X  C  R"  is  said  to  be  compact  if  it  is  both  closed 
and  bounded. 


Recall  that  X  is  closed  if  it  contains  all  the  points  that  make  up  its  boundary. 
(See  Definition  2.3  of  §2.2.)  To  say  that  X  is  bounded  means  that  there  is  some 
(open  or  closed)  ball  B  that  contains  it.  (That  is,  X  is  bounded  if  there  is 
some  positive  number  M  such  that  ||x||  <  M  for  all  x  e  X.)  Thus,  compact  sets 
contain  their  boundaries  (a  consequence  of  being  closed)  and  have  only  finite 
extent  (a  consequence  of  being  bounded).  Some  typical  compact  sets  in  R2  and 
R3  are  shown  in  Figure  4.20. 

For  our  purposes  the  notion  of  compactness  is  of  value  because  of  the  next 
result,  which  we  state  without  proof. 


THEOREM  2.5  (Extreme  value  theorem)  If  X  c  R"  is  compact  and  /: 
X  ->  R  is  continuous,  then  /  must  have  both  a  global  maximum  and  a  global 
minimum  somewhere  on  X.  That  is,  there  must  exist  points  amax  and  am;n  in  X 
such  that,  for  all  x  e  X, 

/(amin)  <  /(x)  <  /(amax). 


We  need  the  compactness  hypothesis  since  a  function  defined  over  a  noncom- 
pact  domain  may  increase  or  decrease  without  bound  and  hence,  fail  to  have  any 
global  extremum,  as  suggested  by  Figure  4.21.  This  is  analogous  to  the  situation 
in  one  variable  where  a  continuous  function  defined  on  an  open  interval  may  fail 
to  have  any  extrema,  but  one  defined  on  a  closed  interval  (which  is  a  compact 
subset  of  R)  must  attain  both  maximum  and  minimum  values.  (See  Figure  4.22.) 
In  the  one-variable  case,  extrema  can  occur  either  in  the  interior  of  the  interval 
or  else  at  the  endpoints.  Therefore,  you  must  compare  the  values  of  /  at  any 
interior  critical  points  with  those  at  the  endpoints  to  determine  which  is  largest 
and  smallest.  In  the  case  of  functions  of  n  variables,  we  do  something  similar, 
namely,  compare  the  values  of  /  at  any  critical  points  with  values  at  any  restricted 
critical  points  that  may  occur  along  the  boundary  of  the  domain. 


272       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


x 


Figure  4.21  A  graph  that  lacks  a 

global  minimum.  Figure  4.22  The  function  depicted  by  the  graph  on  the  left  has  no  global  extrema — the 

function  is  defined  on  the  open  interval  (a,  b).  By  contrast,  the  function  defined  on  the 
closed  interval  [a,  b],  and  with  the  graph  on  the  right,  has  both  a  global  maximum  and 
minimum. 


EXAMPLE  8  LetTiXcR2 


y  =  2 

x  =  -l 

x  =  2 

y  =  -l 

R  be  given  by 


Figure  4.23  The  domain  of  the 
function  T  of  Example  8. 


T(x,  y)  =  x2  -xy  +  y2+  1, 

where  X  is  the  closed  square  in  Figure  4.23.  (Note  that  X  is  compact.)  Think  of 
the  square  as  representing  a  fiat  metal  plate  and  the  function  T  as  the  temperature 
of  the  plate  at  each  point.  Finding  the  global  extrema  amounts  to  finding  the 
warmest  and  coldest  points  on  the  plate.  According  to  Theorem  2.5,  such  points 
must  exist. 

We  need  to  find  all  possible  critical  points  of  T.  Momentarily  considering  T 
as  a  function  on  all  of  R2,  we  find  the  usual  critical  points  by  setting  DT(x,  y) 
equal  to  0.  The  result  is  the  system  of  two  equations 

|    2x  -  y  =  0 
|-x  +  2y  =  0' 

which  has  (0, 0)  as  its  only  solution.  Whether  it  is  a  local  maximum  or  minimum 
is  not  important  for  now,  because  we  seek  global  extrema.  Because  there  is  only 
one  critical  point,  at  least  one  global  extremum  must  occur  along  the  boundary  of 
X  (which  consists  of  the  four  edges  of  the  square).  We  now  find  all  critical  points 
of  the  restriction  of  T  to  this  boundary: 

1.  The  bottom  edge  of  X  is  the  set 

El={(x,y)\y  =  -l,-l<x  <2}. 

The  restriction  of  T  to  E\  defines  a  new  function  f\ :  [—  1 , 2]  — >  R  given  by 

fx(x)=  T(x,-\)  =  x2  +x  +  2. 

As  f{(x)  =  2x  +  1,  the  function  f\  has  a  critical  point  at  x  =  —  j.  Thus, 
we  must  examine  the  following  points  of  X  for  possible  extrema:  (—  \ ,  —  l), 
(—  1 ,  —  1),  and  (2,  —  1).  (The  first  point  is  the  critical  point  of  f\ ,  and  the  second 
two  are  the  vertices  of  X  that  lie  on  E\ .) 

2.  The  top  edge  of  X  is  given  by 


E2  =  {(x,y)  \  y  =  2, -1  <x  <  2}. 


4.2  |  Extrema  of  Functions  273 


Consequently,  we  define  fy.  [—  1 ,  2]  -»  R  by 

/2(jc)  =  T(jr,  2)  =  x2  -  2x  +  5. 

(/2  is  the  restriction  of  71  to  £2-)  We  calculate  f^ipc)  =  2x  —  2,  which  implies 
that  x  =  1  is  a  critical  point  of  /2.  Hence,  we  must  consider  (1,  2),  (—1,  2), 
and  (2,  2)  as  possible  sites  for  global  extrema  of  T.  (The  points  (—1,2)  and 
(2,  2)  are  the  remaining  two  vertices  of  X.) 

3.  The  left  edge  of  X  is 

E3  =  {(x,y)\x  =  -l,-\  <  v<2}. 

Therefore,  we  define  fa  :  [—  1 ,  2]  ->  R  by 

/3(y)=r(-l,y)  =  y2  +  y  +  2. 

We  have  /3'(y)  =  2y  +  1,  and  so  y  =  —  I  is  the  only  critical  point  of  fa.  Thus 
(— 1 ,  —  2)  is  a  potential  site  of  a  global  extremum.  (We  need  not  worry  again 
about  the  vertices  (—1,-1)  and  (—  1 ,  2).) 

4.  The  right  edge  of  X  is 

£4  =  {(x,y)|x=2,-l  <y<2}. 

We  define /4:  [-1,2]  ->  Rby 

f4(y)  =  7X2,  y)  =  y2  -  2y  +  5. 

We  have  /4'(y)  =  2y  —  2,  and  so  y  =  1  is  the  only  critical  point  of  /4.  Hence, 
we  must  include  (2,  1)  in  our  consideration. 

Consequently,  we  have  nine  possible  locations  for  global  extrema,  shown  in 
Figure  4.24.  Now  we  need  only  to  compare  the  actual  values  of  T  at  these  points 
to  see  that  (0,  0)  is  the  coldest  point  on  the  plate  and  both  (2,  —  1)  and  (—1,2)  are 
the  hottest  points.  ♦ 

If  a  function  is  defined  over  a  noncompact  region,  there  is  no  general  result 
like  the  extreme  value  theorem  (Theorem  2.5)  to  guarantee  existence  of  any  global 
extrema.  However,  ad  hoc  arguments  frequently  can  be  used  to  identify  global 
extrema. 


(-1,2) 


(1.2)  (2,2) 


(0,0) 


(-1,-1)' 
(-1,-1)/ 

H,-i) 


(2,1) 


(2,-1) 


(*o0 

T(x,y) 

(0,0) 

1 

(-5,-1) 

7 
4 

(-1,-1) 

2 

(2,-1) 

8 

(1,2) 

4 

(-1,2) 

8 

(2,2) 

5 

(-1,-1) 

7 
4 

(2,1) 

4 

Figure  4.24  Possible  global  extrema  for  T . 


274       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


EXAMPLE  9  Consider  the  function  f(x,  y)  =  e1'3^  denned  on  all  of  R2 
(so  the  domain  is  certainly  not  compact).  Verifying  that  /  has  a  unique  critical 
point  at  (0,  0)  is  straightforward.  We  leave  it  to  you  to  check  that  the  Hessian 
criterion  implies  that  /  has  a  local  maximum  there.  In  any  case,  for  all  (x,y)  6  R2, 
we  have 

1  -  3x2  -  y2  <  1. 

Therefore,  because  the  exponential  function  is  always  increasing  (i.e.,  if  u\  <  u%, 
thene"1  <  e"2), 

As  f(0,  0)  =  e,  we  see  that  /  has  a  global  maximum  at  (0,  0).  ♦ 

WARNING  It  is  tempting  to  assume  that  if  a  function  has  a  unique  critical  point 
that  is  a  local  extremum,  then  it  must  be  a  global  extremum  as  well.  Although 
true  for  the  case  of  a  function  of  a  single  variable,  it  is  not  true  for  functions  of 
two  or  more  variables.  (See  Exercise  52  for  an  example.) 

Addendum:  Proof  of  Theorem  2.3   

Step  1.  We  show  the  following  key  property  of  a  quadratic  form  Q,  namely,  that 
if  k  e  R,  then 

Q(Xh)  =  X2Q(h).  (3) 

This  is  straightforward  to  establish  if  we  write  Q  in  terms  of  its  associated  sym- 
metric matrix  B  and  use  some  of  the  properties  of  matrix  arithmetic  given  in  §  1 .6: 

g(A.h)  =  (Xh)TB(Xh)  =  khTB(Xh)  =  X2hT  Bh  =  A2g(h). 

Step  2.  We  show  that  if  B  is  the  symmetric  matrix  associated  to  a  positive 
definite  quadratic  form  Q,  then  there  is  a  positive  constant  M  such  that 

e(h)  >  Mpii2 

for  allh  6  R". 

First,  note  that  when  h  =  0,  then  Q(h)  =  <2(0)  =  0  so  the  conclusion  holds 
trivially  in  this  case. 

Next,  suppose  that  his  a  unit  vector  (i.e.,  ||h||  =  1).  The  (endpoints  of  the)  set 
of  all  unit  vectors  in  R"  is  an  (n  —  l)-dimensional  sphere  S,  which  is  a  compact 
set.  Hence,  by  the  extreme  value  theorem  (Theorem  2.5),  the  restriction  of  Q  to 
S  must  achieve  a  global  minimum  value  M  somewhere  on  S.  Thus,  Q(h)  >  M 
for  all  h  €  S. 

Finally,  let  h  be  any  nonzero  vector  in  R".  Then  its  normalization  h/||h||  is  a 
unit  vector  and  so  lies  in  S.  Therefore,  by  the  result  of  Step  1,  we  have 

G(h)=e(l|l.|l1|||)  =  l|hfG(1|I)>IIN2«, 

since  h/||h||  is  in  S. 

Step  3.  Now  we  prove  the  theorem.  By  the  second-order  Taylor  formula 
Theorem  1.5  and  formula  (10)  of  §4.1,  we  have  that,  for  the  critical  point  a  of  /, 

Af  =  f(x)  -  /(a)  =  i(x  -  a)r///(a)(x  -  a)  +  R2(x,  a),  (4) 
where  |i?2(x,  a)|/||x  —  a||2  ->  0  as  x  -»  a. 


4.2  |  Extrema  of  Functions  275 


Suppose  first  that  Hf(&)  is  positive  definite.  Then  by  Step  2  with  h  =  x  —  a, 
there  must  exist  a  constant  M  >  0  such  that 

i(x-a)r///(a)(x-a)>  M||x-a||2.  (5) 

Because  |i?2(x,  a)|/||x  —  a||2  ->  0  as  x  ->  a,  there  must  be  some  5  >  0  so  that  if 
0  <  ||x  -  a||  <  <5,  then  |i?2(x,  a)|/||x  -  a||2  <  M,  or,  equivalently, 

|i?2(x,a)|  <M||x-a||2.  (6) 

Therefore,  (4),  (5),  and  (6)  imply  that,  for  0  <  ||x  —  a||  <  <5, 

A/>0 

so  that  /  has  a  (strict)  local  minimum  at  a. 

If  Hf(a)  is  negative  definite,  then  consider  g  =  —f.  We  see  that  a  is  also  a 
critical  point  of  g  and  that  Hg(a)  =  —  Hf(a),  so  Hg(a)  is  positive  definite.  Hence, 
the  argument  in  the  preceding  paragraph  shows  that  g  has  a  local  minimum  at  a, 
so  /  has  a  local  maximum  at  a. 

Now  suppose  det  Hf(a)  0,  but  that  Hf(a)  is  neither  positive  nor  negative 
definite.  Let  xi  be  such  that 

i(x,  -  a)r///(a)(Xl  -  a)  >  0 

and  X2  such  that 

i(x2  -  a)r///(a)(x2  -  a)  <  0. 

(Since  det  Hf(a)     0,  such  points  must  exist.)  For  i  =  1,2  let 

yt(0  =  t(x  -  a)  +  a, 

the  vector  parametric  equation  for  the  line  through  a  and  x,- .  Applying  formula 
(4)  with  x  =  y,-(f),  we  see 

A/  =  f(Yi(t))  -  f(a)  =  i(y,(f)  -  a)rH/(a)(y,(0  "  a)  +  ^(y,(0,  a) 
=  i(y(-  0  -  a)rJf/(aXy,(0  -  a)  +  ||y,(f)  -  a||2     g  / 

lly/(0-a||2 

Note  that  y,  (f)  —  a  =  f  (x,  —  a).  Therefore,  using  the  property  of  quadratic  forms 
given  in  Step  1  and  the  fact  that  ||y,(?)  -  a||2  =  ||?(x;  -  a)||2  =  r2||x/  -  a||2,  we 
have 

/(y  (0)  -  /(a) 

(7) 


=  t2 


i(x;  -  tfHfW*  -  a)  +  ||x,  -  a||2  f  2^(0'  ^ 

llyi(0  -  all2 


Now  note  that,  for  i  =  1 ,  the  first  term  in  the  brackets  in  the  right  side  of  (7)  is  a 
positive  number  P  and,  for  i  =  2,  it  is  a  negative  number  N.  Set 


M  =  min 


|X!  -  a||2 '  ||x2-a||2 


Because  we  know  that  \Rz(yi(t),  a)|/||y(r)  -  a||2  ->  0  as  t  -+  0,  we  can  find 
some  S  >  0  so  that  if  0  <  t  <  S,  then 

1*2(7,(0,  a)|  <M 


|y(0  -a||5 


276       Chapter  4  [  Maxima  and  Minima  in  Several  Variables 


But  this  implies  that,  for  0  <  t  <  S, 

A/  =  /(yi(O)-/(a)>0, 

while 

A/  =  /(y2(0)  "  /(*)  <  0. 
Thus,  /  has  a  saddle  point  at  x  =  a. 


4.2  Exercises 


1.  Concerning  the  function  f(x,  y)  =  Ax  +  6y  —  12  — 

2  2 

x  —  y  : 

(a)  There  is  a  unique  critical  point.  Find  it. 

(b)  By  considering  the  increment  Af,  determine 
whether  this  critical  point  is  a  maximum,  a  mini- 
mum, or  a  saddle  point. 

(c)  Now  use  the  Hessian  criterion  to  determine  the 
nature  of  the  critical  point. 

2.  This  problem  concerns  the  function  g(x,  y)  =  x2  — 
2y2  +  2x  +  3. 

(a)  Find  any  critical  points  of  g. 

(b)  Use  the  increment  Ag  to  determine  the  nature  of 
the  critical  points  of  g. 

(c)  Use  the  Hessian  criterion  to  determine  the  nature 
of  the  critical  points. 

In  Exercises  3-20,  identify  and  determine  the  nature  of  the 
critical  points  of  the  given  functions. 


3. 

/(*.  y)  = 

2xy  -  2x2  -  5y2  +  Ay  -  3 

4. 

fix,  y)  = 

ln(x2  +  y2+  1) 

5. 

fix,  y)  = 

x2  +  y3  —  6xy  +  3x  +  6y 

6. 

fix,  y)  = 

y4  —  2xy2  +  x3  —  x 

7. 

fix,  y)  = 

8  1 

XV  +  -  +  - 

x  y 

8. 

fix,  y)  = 

ex  sin  y 

9. 

fix,  y)  = 

e-y(x2  -  y2) 

10. 

fix,  y)  = 

{X  +  y)(l  -  xy) 

11. 

fix,  y)  = 

x  —  y  —  x  y  +  y 

12. 

fix,y)  = 

e~x(x2  +  3y2) 

13. 

fix,  y)  = 

2x  —  3y  +  Inxy 

14. 

fix,  y)  = 

cosx  siny 

15. 

fix,y,z) 

=  x2  —  xy  +  z2  —  2xz  +  6z 

16. 

fix,  y,  z) 

=  (x2+2y2  +  l)cosz 

17. 

fix,  y,  z) 

=  x2  +  y2  +  2z2  +  xz 

18.  f(x,  y,  z)  =  jc3  +  xz2  -  3x2  +  y2  +  2z2 

1 

19.  f(x,y,z)  =  xy  +  xz  +  2yz+  - 

x 

20.  f(x,y,z)  =  ex(x2  -  y2  -  2z2) 

21.  (a)  Find     all     critical     points     of     fix,y)  = 

2y3  -  3y2  -  36y  +  2 
1  +  3x2  ' 
(b)  Identify  any  and  all  extrema  of  /. 

22.  (a)  Under  what  conditions  on  the  constant  k  will  the 

function 

f{x,  y)  =  kx2  —  2xy  +  ky2 

have  a  nondegenerate  local  minimum  at  (0,  0)? 
What  about  a  local  maximum? 

(b)  Under  what  conditions  on  the  constant  k  will  the 
function 

g(x,  y,  z)  =  kx2  +  kxz  —  2yz  —  y2  +  -Z2 

have  a  nondegenerate  local  maximum  at  (0,  0,  0)? 
What  about  a  nondegenerate  local  minimum? 

23.  (a)  Consider    the    function    f(x,  y)  =  ax2  +  by2, 

where  a  and  b  are  nonzero  constants.  Show  that 
the  origin  is  the  only  critical  point  of  /,  and  deter- 
mine the  nature  of  that  critical  point  in  terms  of  a 
and  b. 

(b)  Now  consider  the  function  f(x,y,z)  =  ax2  + 
by2  +  cz2,  where  a,  b,  and  c  are  all  nonzero.  Show 
that  the  origin  in  R3  is  the  only  critical  point  of  /, 
and  determine  the  nature  of  that  critical  point  in 
terms  of  a,  b,  and  c. 

(c)  Finally,  let  fix\,  x%,  x„)  =  a\x\  +  02X2 
+  ■  ■  ■  +  a,,x2,  where  a;  is  a  nonzero  constant  for 
i  =  1,  2, . . . ,  n.  Show  that  the  origin  in  R"  is  the 
only  critical  point  of  /,  and  determine  its  nature. 

Sometimes  it  can  be  difficult  to  determine  the  critical  point  of 
a  function  f  because  the  system  of  equations  that  arises  from 
setting  V/  equal  to  zero  may  be  very  complicated  to  solve  by 
hand.  For  the  functions  given  in  Exercises  24—27,  (a)  use  a 
computer  to  assist  you  in  identifying  all  the  critical  points  of 
the  given  function  f ,  and  (b)  use  a  computer  to  construct  the 


4.2  I  Exercises  277 


Hessian  matrix  and  determine  the  nature  of  the  critical  points 
found  in  part  (a). 

^  24.  f(x,  y)  =  y4  +  x3  -  2xy2  -  x 
^>  25.  f(x,  y)  =  2x3y  -  y2  -  3xy 

^  26.  f(x,  y,  z)  =  yz-  xyz  -  x2  -  y2  -  2z2 

27.  f(x,  y,  z,  u>)  =  yw  —  xyz  —  x2  —  2z2  +  w2 

28.  Show  that  the  largest  rectangular  box  having  a  fixed 
surface  area  must  be  a  cube. 

29.  What  point  on  the  plane  3x  —  4y  —  z  =  24  is  closest 
to  the  origin? 

30.  Find  the  points  on  the  surface  xy  +  z2  =  4  that  are 
closest  to  the  origin.  Be  sure  to  give  a  convincing  ar- 
gument that  your  answer  is  correct. 

31 .  Suppose  that  you  are  in  charge  of  manufacturing  two 
types  of  television  sets.  The  revenue  function,  in  dol- 
lars, is  given  by 

R(x,  y)  =  8x  +  6y  -  x2  -  2y2  +  2xy, 

where  x  denotes  the  quantity  of  model  X  sets  sold,  and 
y  the  quantity  of  model  Y  sets  sold,  both  in  units  of 
100.  Determine  the  quantity  of  each  type  of  set  that 
you  should  produce  in  order  to  maximize  the  resulting 
revenue. 

32.  Find  the  absolute  extrema  of  f(x,  y)  =  x2  +  xy  + 
y2  —  6y  on  the  rectangle  {(x,  y)  \  —  3  <  x  <  3,  0  < 

y  <  5}'. 

33.  Find  the  absolute  maximum  and  minimum  of 

fix,  y,  z)  =  x2  +  xz  —  y2  +  2z2  +  xy  +  5x 

on  the  block  {(x,  y,z)\  -  5  <  x  <  0,  0  <  y  <  3, 
0  <  z  <  2}. 

34.  A  metal  plate  has  the  shape  of  the  region  x2  +  y2  <  1 . 
The  plate  is  heated  so  that  the  temperature  at  any  point 
(x ,  y)  on  it  is  indicated  by 

T(x,y)  =  2x2  +  y2-y  +  3. 

Find  the  hottest  and  coldest  points  on  the  plate  and  the 
temperature  at  each  of  these  points.  (Hint:  Parametrize 
the  boundary  of  the  plate  in  order  to  find  any  critical 
points  there.) 

35.  Find  the  (absolute)  maximum  and  minimum  values  of 
f(x,  y)  =  sinx  cosy  on  the  square  R  =  {(x,  y)  |  0  < 
x  <  2it,  0  <  y  <  2jt}. 

36.  Find  the  absolute  extrema  of  f(x,  y)  =  2  cosx  + 
3  siny  on  the  rectangle  {(x,  y)  \  0  <  x  <  4,  0  < 

y  <3}. 

37.  Determine  the  absolute  minimum  and  maximum 
values  of  the  function  f(x,  y)  =  2x2  —  2xy  +  y2 


—  y  +  3  on  the  closed  triangular  region  with  vertices 
(0,  0),  (2,  0),  and  (0,  2). 

38.  Determine  the  absolute  minimum  and  maximum  val- 
ues of  the  function  f(x,  y)  =  x2y  on  the  elliptical  re- 
gion D  =  {(x,  y)  |  3x2  +  4y2  <  12}. 

39.  Find  the  absolute  extrema  of  f(x,  y,  z)  = 
e!-x2_y2+2y_z2_4z  Qn  ^  ball  {(^  _^  z)  |  x2  +  y2  -  2y 

+  z2  +  4z  <  0}. 

Each  of  the  functions  in  Exercises  40-45  has  a  critical  point 
at  the  origin.  For  each  function,  (a)  check  that  the  Hessian 
fails  to  provide  any  information  about  the  nature  of  the  critical 
point  at  the  origin,  and  (b)  find  another  way  to  determine  if  the 
function  has  a  maximum,  minimum,  or  neither  at  the  origin. 

40.  f(x,y)  =  x2y2 

41.  f(x,y)  =  4-3x2y2 

42.  f(x,y)  =  x3y3 

43.  f(x,  y,  z)  =  x2y3z4 

44.  f(x,  y,  z)  =  x2y2z4 

45.  f(x,y,z)  =  2-x4y4  -z4 

In  Exercises  46—48,  (a)  find  all  critical  points  of  the  given 
function  f  and  identify  their  nature  as  local  extrema  and  (b) 
determine,  with  explanation,  any  global  extrema  of  f. 

46.  f(x,y)  =  ex2+5>2 

47.  f(x,y,  z)  =  e2-*2-2}2-3'-4 

48.  f(x,  y)  =  x3  +  y3  -  3xy  +  7 

49.  Determine  the  global  extrema,  if  any,  of 

f(x,  y)  =  xy  +  2y  —  lnx  —  2 lny, 
where  x,  y  >  0. 

50.  Find  all  local  and  global  extrema  of  the  function 

f{x,  y,  z)  =  x3  +  3x2  +  e>,2+l  +z2-  3xz. 

51.  Let  f(x,  y)  =  3  -  [(x  -  l)(y  -  2)]2/3. 

(a)  Determine  all  critical  points  of  /. 

(b)  Identify  all  extrema  of  /. 

52.  (a)  Suppose  /:  R  — >  R  is  a  differentiate  function  of 

a  single  variable.  Show  that  if  /  has  a  unique  crit- 
ical point  at  xo  that  is  the  site  of  a  strict  local  ex- 
tremum  of  /,  then  /  must  attain  a  global  extremum 
at  xo. 

(b)  Let  fix,  y)  =  3yex  -  e3x  -  y3.  Verify  that  /  has 
a  unique  critical  point  and  that  /  attains  a  local 
maximum  there.  However,  show  that  /  does  not 
have  a  global  maximum  by  considering  how  /  be- 
haves along  the  y-axis.  Hence,  the  result  of  part 
(a)  does  not  carry  over  to  functions  of  more  than 
one  variable. 


278       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


53.  (a)  Let  /  be  a  continuous  function  of  one  variable. 
Show  that  if  /  has  two  local  maxima,  then  /  must 
also  have  a  local  minimum. 

(b)  The  analogue  of  part  (a)  does  not  necessarily  hold 
for  continuous  functions  of  more  than  one  variable, 


as  we  now  see.  Consider  the  function 


f(x,y)  =  2-(xy2 


D2-(v2-l)2 


Show  that  /  has  just  two  critical  points — and  that 
both  of  them  are  local  maxima. 

(c)  Use  a  computer  to  graph  the  function  /  in  part  (b). 


4.3  Lagrange  Multipliers 

Constrained  Extrema  

Frequently,  when  working  with  applications  of  calculus,  you  will  find  that  you 
do  not  need  simply  to  maximize  or  minimize  a  function  but  that  you  must  do  so 
subject  to  one  or  more  additional  constraints  that  depend  on  the  specifics  of  the 
situation.  The  following  example  is  a  typical  situation: 

EXAMPLE  1  An  open  rectangular  box  is  to  be  manufactured  having  a  (fixed) 
volume  of  4  ft3 .  What  dimensions  should  the  box  have  so  as  to  minimize  the 
amount  of  material  used  to  make  it? 

We'll  let  the  three  dimensions  of  the  box  be  independent  variables  x,  y,  and 
z,  shown  in  Figure  4.25.  To  determine  how  to  use  as  little  material  as  possible, 
we  need  to  minimize  the  surface  area  function  A  given  by 

A(x,  y,  z)  =      2xy      +  2yz  +  xz 

front  and  back       sides       bottom  only 

For  x,  y,  z  >  0,  this  function  has  neither  minimum  nor  maximum.  However,  we 
have  not  yet  made  use  of  the  fact  that  the  volume  is  to  be  maintained  at  a  constant 
4  ft3 .  This  fact  provides  a  constraint  equation, 

V(x,  y,  z)  =  xyz  =  4. 

The  constraint  is  absolutely  essential  if  we  are  to  solve  the  problem.  In  particular, 
the  constraint  enables  us  to  solve  for  z  in  terms  of  x  and  y: 

4 

xy' 

We  can  thus  create  a  new  area  function  of  only  two  variables: 

/  4 

a(x,  y)  =  A  j  x,  y,  — 
V  xy 

8  4 
=  2xy+  -  +  -. 

x  y 

Now  we  can  find  the  critical  points  of  a  by  setting  Da  equal  to  0: 


v 


Figure  4.25  The  open  box  of 
Example  1. 


4.3  |  Lagrange  Multipliers 


The  first  equation  implies 


so  that  the  second  equation  becomes 


2x  -  4 


16 


or,  equivalently, 


x    1  -  -x*    =  0 


The  solutions  to  this  equation  are  x  =  0  (which  we  reject)  and  x  =  2.  Thus,  the 
critical  point  of  a  of  interest  is  (2,  1),  and  the  constrained  critical  point  of  the 
original  function  A  is  (2,  1,2). 

We  can  use  the  Hessian  criterion  to  check  that  x  =  2,  y  =  1  yields  a  local 
minimum  of  a : 


Ha(x,  y)  = 


16/x3 
2 


2 

8/y3 


so    Ha(2,  1)  = 


The  sequence  of  minors  is  2,  12  so  we  conclude  that  (2,  1)  does  yield  a  local 
minimum  of  a.  Because  a(x,  y)  — >■  oo  as  either  jc  — >  0+,  y  — >  0+,  je  — >  oo,  or 
y  ->  oo,  we  conclude  that  the  critical  point  must  yield  a  global  minimum  as  well. 
Thus,  the  solution  to  the  original  question  is  to  make  the  box  with  a  square  base 
of  side  2  ft  and  a  height  of  1  ft.  ♦ 


The  abstract  setting  for  the  situation  discussed  in  Example  1  is  to  find  max- 
ima or  minima  of  a  function  f(x\,  xi,  ■  ■  ■ ,  xn)  subject  to  the  constraint  that 
g{x\,  X2,  ■  ■  ■ ,  xn)  =  c  for  some  function  g  and  constant  c.  (In  Example  1,  the 
function  /  is  A(x,  y,  z),  and  the  constraint  is  xyz  =  4.)  One  method  for  finding 
constrained  critical  points  is  used  implicitly  in  Example  1:  Use  the  constraint 
equation  g(x)  =  c  to  solve  for  one  of  the  variables  in  terms  of  the  others.  Then 
substitute  for  this  variable  in  the  expression  for  /(x),  thereby  creating  a  new 
function  of  one  fewer  variables.  This  new  function  can  then  be  maximized  or 
minimized  using  the  techniques  of  §4.2.  In  theory,  this  is  an  entirely  appropriate 
way  to  approach  such  problems,  but  in  practice  there  is  one  major  drawback:  It 
may  be  impossible  to  solve  explicitly  for  any  one  of  the  variables  in  terms  of  the 
others.  For  example,  you  might  wish  to  maximize 

f(x,y,z)  =  x2  +  3y2  +  y2z4 


subject  to 


g(x,y,z)  =  ex 


x5y2z  +  cos 


=  2. 


There  is  no  means  of  isolating  any  of  x,  y,  or  z  on  one  side  of  the  constraint 
equation,  and  so  it  is  impossible  for  us  to  proceed  any  further  along  the  lines  of 
Example  1. 


280       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


The  Lagrange  Multiplier  

The  previous  discussion  points  to  the  desirability  of  having  another  method  for 
solving  constrained  optimization  problems.  The  key  to  such  an  alternative  method 
is  the  following  theorem: 


THEOREM  3.1  Let  X  be  open  in  R"  and  f,g:X^Rbe  functions  of  class  C1 . 
Let  S  =  {x  e  X  \  g(x)  =  c}  denote  the  level  set  of  g  at  height  c.  Then  if  /  \s  (the 
restriction  of  /  to  S)  has  an  externum  at  a  point  x0eS  such  that  Vg(x0)  0, 
there  must  be  some  scalar  X  such  that 

V/(x0)  =  AVg(x0). 


The  conclusion  of  Theorem  3.1  implies  that  to  find  possible  sites  for  extrema 
of  /  subject  to  the  constraint  that  g(x)  =  c,  we  can  proceed  in  the  following 
manner: 

1.  Form  the  vector  equation  V/(x)  =  XVg(x). 

2.  Solve  the  system 

V/(x)  =  AVg(x) 
g(X)  =  c 

for  x  and  X.  When  expanded  this  is  actually  a  system  of  n  +  1  equations  in 
n  +  1  unknowns  x\,  x2,  ■  ■  ■ ,  xn,  X,  namely, 

fXl(xi,x2,  ■  ■  .,x„)  =  kgXl(xi,x2,  ■  ■  -,x„) 
fX2(xi,x2,  ...,xn)  =  XgX2(xi,x2,  ■  ■  ■ ,  xn) 

• 

fx„(xi  >x2,...,xn)  =  XgXiXxi  ,x2, xn) 
g(xux2,  . . .  ,x„)  =  c 

The  solutions  for  x  =  (x\,  x2, . . . ,  xn)  in  the  system  above,  along  with  any 
other  points  x  satisfying  the  constraint  g(x)  =  c  and  such  that  V/  is  unde- 
fined or  Vg  vanishes  or  is  undefined,  are  the  candidates  for  extrema  for  the 
problem. 

3.  Determine  the  nature  of  /  (as  maximum,  minimum,  or  neither)  at  the  critical 
points  found  in  Step  2. 

The  scalar  A  appearing  in  Theorem  3. 1  is  called  a  Lagrange  multiplier,  after 
the  Italian-born  French  mathematician  Joseph-Louis  Lagrange  (1736-1813)  who 
first  developed  this  method  for  solving  constrained  optimization  problems.  In 
practice,  Step  2  can  involve  some  algebra,  so  it  is  important  to  keep  your  work 
organized.  (Alternatively,  you  can  use  a  computer  to  solve  the  system.)  In  fact, 
since  the  Lagrange  multiplier  X  is  usually  not  of  primary  interest,  you  can  avoid 
solving  for  it  explicitly,  thereby  reducing  the  algebra  and  arithmetic  somewhat. 
Determining  the  nature  of  a  constrained  critical  point  (Step  3)  can  be  a  tricky 
business.  We'll  have  more  to  say  about  that  issue  in  the  examples  and  discussions 
that  follow. 

EXAMPLE  2  Let  us  use  the  method  of  Lagrange  multipliers  to  identify  the 
critical  point  found  in  Example  1 .  Thus,  we  wish  to  find  the  minimum  of 

A(x,  y,  z)  =  2xy  +  2yz  +  xz 


4.3  |  Lagrange  Multipliers  281 


subject  to  the  constraint 

V(x,  y,  z)  =  xyz  =  4. 
Theorem  3.1  suggests  that  we  form  the  equation 

VA(x,y,z)  =  XVV(x,y,z). 

This  relation  of  gradients  coupled  with  the  constraint  equation  gives  rise  to  the 
system 

2y  +     z  =  kyz 
2x  +   2z  =  ~kxz 
2y  +     x  =  kxy 
xyz  =  4 

Since  A  is  not  essential  for  our  final  solution,  we  can  eliminate  it  by  means  of  any 
of  the  first  three  equations.  Hence, 

_  2y  +  z  _2x  +2z  _2y  +  x 

yz  xz  xy 

Simplifying,  this  implies  that 

2      1      2     2      2  1 
-  +  -  =  -  +  -  =  -  +  -. 

z     y      z     x      x  y 

The  first  equality  yields 

—  =  —    or    x  =  2y, 

y  x 

while  the  second  equality  implies  that 

2  1 

-  =  -    or    z  =  2y. 

z  y 

Substituting  these  relations  into  the  constraint  equation  xyz  =  4  yields 

(2y)(y)(2y)  =  4, 

so  that  we  find  that  the  only  solution  isy  =  l,je=z  =  2,  which  agrees  with  our 
work  in  Example  1.  (Note  that  V  V  =  0  only  along  the  coordinate  axes,  and  such 
points  do  not  satisfy  the  constraint  V(x,  y,  z)  =  4.)  ♦ 


An  interesting  consequence  of  Theorem  3. 1  is  this:  By  Theorem  6.4  of  Chap- 
ter 2,  we  know  that  the  gradient  Vg,  when  nonzero,  is  perpendicular  to  the  level 
sets  of  g.  Thus,  the  equation  V/  =  XV g  gives  the  condition  for  the  normal  vector 
to  a  level  set  of  /  to  be  parallel  to  that  of  a  level  set  of  g.  Hence,  for  a  point  xo 
to  be  the  site  of  an  extremum  of  /'  on  the  level  set  S  =  {x  |  g(x)  =  c],  where 
Vg(x0)  ^  0,  we  must  have  that  the  level  set  R  of  /  that  contains  x0  is  tangent  to 
S  at  xo. 

EXAMPLE  3  Consider  the  problem  of  finding  the  extrema  of  f(x,  y)  =  x2/4 
+  y2  subject  to  the  condition  that  x2  +  y2  =  1.  We  let  g(x,  y)  =  x2  +  y2,  and 
so  the  Lagrange  multiplier  equation  V  f(x,  y)  =  XVg(x,  y),  along  with  the 


282       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


constraint  equation,  yields  the  system 

x 


Figure  4.26  The  level  sets  of  the 
function  f(x,  y)  =  x2/4  +  y2 
define  a  family  of  ellipses.  The 
extrema  of  /  subject  to  the 
constraint  that  x2  +  y2  =  1  (i.e., 
that  lie  on  the  unit  circle)  occur  at 
points  where  an  ellipse  of  the 
family  is  tangent  to  the  unit  circle. 


=  2Xx 


2y  =  2Xy  ■ 

x2  +  y2  =  1 

(There  are  no  points  simultaneously  satisfying  g(x,  y)  =  1  and  Vg(x,y)  = 
(0,  0).)  The  first  equation  of  this  system  implies  that  either  x  =  0  or  X  =  \. 
If  x  =  0,  then  the  second  two  equations,  taken  together,  imply  that  y  =  ±1  and 
X  =  1 .  If  X  =  \,  then  the  second  two  equations  imply  y  =  0  and  x  —  ±1 .  There- 
fore, there  are  four  constrained  critical  points:  (0,  ±1),  corresponding  to  X  =  1, 
and  (±1,  0),  corresponding  to  X  =  |. 

We  can  understand  the  nature  of  these  critical  points  by  using  geometry  and 
the  preceding  remarks.  The  collection  of  level  sets  of  the  function  /  is  the  family 
of  ellipses  x2/4  +  y2  =  k  whose  major  and  minor  axes  lie  along  the  x-  and  y- 
axes,  respectively.  In  fact,  the  value  f(x,  y)  =  x2 /A  +  y2  =  k  is  the  square  of 
the  length  of  the  semiminor  axis  of  the  ellipse  x2/4  +  y2  =  k.  The  optimization 
problem  then  is  to  find  those  points  on  the  unit  circle  x2  +  y2  =  1  that,  when 
considered  as  points  in  the  family  of  ellipses,  minimize  and  maximize  the  length 
of  the  minor  axis.  When  we  view  the  problem  in  this  way,  we  see  that  such  points 
must  occur  where  the  circle  is  tangent  to  one  of  the  ellipses  in  the  family.  A  sketch 
shows  that  constrained  minima  of  /  occur  at  (±  1 ,  0)  and  constrained  maxima  at 
(0,  ±  1 ).  In  this  case,  the  Lagrange  multiplier  X  represents  the  square  of  the  length 
of  the  semiminor  axis.  (See  Figure  4.26.)  ♦ 

EXAMPLE  4  Consider  the  problem  of  determining  the  extrema  of  f(x,  y)  = 
2x  +  y  subject  to  the  constraint  that  *Jx  +  Jy  =  3.  We  let  g(x,  y)  =  *Jx  +  ^/y, 
so  that  the  Lagrange  multiplier  equation  V  f(x,  y)  =  XVg(x,  y),  along  with  the 
constraint  equation,  yields  the  system 

X 


2jx 
X 

1  =    ■ 

yfx  +  s/y  =  3 

The  first  two  equations  of  this  system  imply  that  X  =  4^/x  =  2Jy  so  that  Jy  = 
2yfx.  Using  this  in  the  last  equation,  we  find  that  3^/x  =  3  and  hence,  x  —  1. 
Thus,  the  system  of  equations  above  yields  the  unique  solution  (1,4). 

Since  the  constraint  defines  a  closed  bounded  curve  segment,  the  extreme 
value  theorem  (Theorem  2.5)  applies  to  guarantee  that  /  must  attain  both  a 
global  maximum  and  a  global  minimum  on  this  segment.  However,  the  Lagrange 
multiplier  method  has  provided  us  with  just  a  single  critical  point.  But  note  that  the 
points  (9,  0)  and  (0,  9)  satisfy  the  constraint  *Jx  +  ^/y  =  3;  they  are  both  points 
where  Vg  is  undefined.  Moreover,  we  have  /(l,  4)  =  2,  while  f(9,  0)  =  18  and 
f(0,  9)  =  9.  Evidently  then,  the  minimum  of  /  occurs  at  (1 ,  4)  and  the  maximum 
at  (9,0). 

We  can  understand  the  geometry  of  the  situation  in  the  following  manner.  The 
collection  of  level  sets  of  the  function  /  is  the  family  of  parallel  lines  2x  +  y  =  k. 
Note  that  the  height  k  of  each  level  set  is  just  the  y-intercept  of  the  corresponding 
line  in  the  family.  Thus,  the  problem  we  are  considering  is  to  find  the  largest  and 


4.3  |  Lagrange  Multipliers  283 


y 


Figure  4.27  The  level  sets  of  the  function  f(x,  y)  =  2x  +  y  define  a  family 
of  lines.  The  minimum  of  /  subject  to  the  constraint  that  *Jx  +  *Jy  =  3 
occurs  at  a  point  where  one  of  the  lines  is  tangent  to  the  constraint  curve  and 
the  maximum  at  one  of  the  endpoints  of  the  curve. 


smallest  y -intercepts  of  any  line  in  the  family  that  meets  the  curve  *Jx  +  Jy  =  3. 
These  extreme  values  of  k  occur  either  when  one  of  the  lines  is  tangent  to  the 
constraint  curve  or  at  an  endpoint  of  the  curve.  (See  Figure  4.27.) 

This  example  illustrates  the  importance  of  locating  all  the  points  where  ex- 
trema  may  occur  by  considering  places  where  V/  or  Vg  is  undefined  (or  where 
Vg  =  0)  as  well  as  the  solutions  to  the  system  of  equations  determined  using 
Lagrange  multipliers.  ♦ 


Figure  4.28  The  gradient 
Vg(xo)  is  perpendicular  to 
S  =  {x  |  g(x)  =  c},  hence,  to  the 
tangent  vector  at  xo  to  any  curve 
x(r)  lying  in  S  and  passing  through 
xo.  If  /  has  an  extremum  at  xo, 
then  the  restriction  of  /  to  the 
curve  also  has  an  extremum  at  xq. 


Sketch  of  a  proof  of  Theorem  3.1  We  present  the  key  ideas  of  the  proof,  which 
are  geometric  in  nature.  Try  to  visualize  the  situation  for  the  case  n  =  3,  where 
the  constraint  equation  g(x,  y,  z)  =  c  defines  a  surface  S  inR3.  (See  Figure  4.28.) 
In  general,  if  S  is  defined  as  {x  |  g(x)  =  c)  with  Vg(xo)  ^  0,  then  (at  least  locally 
near  xo)  S  is  a  hypersurface  in  R".  The  proof  that  this  is  the  case  involves  the 
implicit  function  theorem  (Theorem  6.5  in  §2.6),  and  this  is  why  our  proof  here 
is  just  a  sketch. 

Thus,  suppose  that  x0  is  an  extremum  of  /  restricted  to  S.  We  consider  a 
further  restriction  of  / — to  a  curve  lying  in  S  and  passing  through  x0.  This  will 
enable  us  to  use  results  from  one-variable  calculus.  The  notation  and  analytic 
particulars  are  as  follows:  Let  x:  /  c  R  —>  S  C  R3  be  a  C1  path  lying  in  S  with 
\(to)  =  xo  for  some  to  6  / .  Then  the  restriction  of  /  to  x  is  given  by  the  function 
F,  where 

F(f)  =  /(x(0). 

Because  xo  is  an  extremum  of  /  on  S,  it  must  also  be  an  extremum  on  x.  Conse- 
quently, we  must  have  F'(to)  =  0,  and  the  chain  rule  implies  that 


=  ^/(x(0)|r=ro  =  V/(x(f0)).x'(fo)  =  V/(x0)-x'(f0). 


Thus,  V/(xo)  is  perpendicular  to  any  curve  in  S  passing  through  x0;  that  is, 
V /(xo)  is  normal  to  S  at  xo.  We've  seen  previously  in  §2.6  that  the  gradient 
Vg(xo)  is  also  normal  to  S  at  xq.  Since  the  normal  direction  to  the  level  set  S  is 


284       Chapter  4  [  Maxima  and  Minima  in  Several  Variables 


uniquely  determined  and  Vg(xo)  7^  0,  we  must  conclude  that  V /(xo)  and  Vg(xo) 
are  parallel  vectors.  Therefore, 

V/(x0)  =  AVg(xo) 

for  some  scalar  X  e  R,  as  desired.  ■ 

The  Case  of  More  than  One  Constraint  

It  is  natural  to  generalize  the  situation  of  finding  extrema  of  a  function  /  subject 
to  a  single  constraint  equation  to  that  of  finding  extrema  subject  to  several  con- 
straints. In  other  words,  we  may  wish  to  maximize  or  minimize  /  subject  to  k 
simultaneous  conditions  of  the  form 

gi(x)  =  c\ 

g2(x)  =  C2 
gk(x)  =  Ck 

The  result  that  generalizes  Theorem  3.1  is  as  follows: 


THEOREM  3.2  Let  X  be  open  in  R"  and  let  /,  gu  . . . ,  gk:  X  C  R"  ->  R  be  C1 
functions, where k  <  n.LetS  =  [x  e  X  \  g\(x)  =  c\, .. .,  gfc(x)  =  ck}.lf  f\shas 
an  extremum  at  a  point  xo,  where  Vgi(xo), . . . ,  Vgt(xo)  are  linearly  independent 
vectors,  then  there  must  exist  scalars  X\, ...  ,Xk  such  that 

V/(xo)  =  AjVgKxo)  +  A2Vg2(x0)  +  ■  ■  ■  +  XkVgk(x0). 

(Note:  k  vectors  vi , . . . ,  in  R"  are  said  to  be  linearly  independent  if  the  only 
way  to  satisfy  a\ vi  +  ■  ■  ■  +  akvk  =  0  for  scalars  a\, . . . ,  ak  is  if  a\  =  a%  =  ■  ■  ■  = 
ak  =  0.) 


Idea  of  proof  First,  note  that  S  is  the  intersection  of  the  k  hypersurfaces  Si , . . . , 
S/c,  where  Sj  =  {x  6  R"  |  g/(x)  =  cj}.  Therefore,  any  vector  tangent  to  S  must 
also  be  tangent  to  each  of  these  hypersurfaces,  and  so,  by  Theorem  6.4  of 
Chapter  2,  perpendicular  to  each  of  the  Vg;  's.  Given  these  remarks,  the  main 
ideas  of  the  proof  of  Theorem  3.1  can  be  readily  adapted  to  provide  a  proof  of 
Theorem  3.2. 

Therefore,  we  let  xo  e  5  be  an  extremum  of  /  restricted  to  S  and  consider 
the  one-variable  function  obtained  by  further  restricting  /  to  a  curve  in  S  through 
xo.  Thus,  let  x:  /  ->  S  C  R"  be  a  C1  curve  in  S  with  x(f0)  =  xo  for  some  to  £  I. 
Then,  as  in  the  proof  of  Theorem  3.1,  we  define  F  by 

F(t)  =  /(x(0). 

It  follows,  since  x0  is  assumed  to  be  a  constrained  extremum,  that 

F'(t0)  =  0. 

The  chain  rule  then  tells  us  that 

0  =  F'(to)  =  V/(x(*0))  •  x'(to)  =  V/(xo)  •  x'(to). 

That  is,  V /(xo)  is  perpendicular  to  all  vectors  tangent  to  S  at  xo.  Therefore,  it  can 
be  shown  that  V/(xq)  is  in  the  ^-dimensional  plane  spanned  by  the  normal  vectors 


4.3  |  Lagrange  Multipliers  285 


Plane  spanned  by 
Vg^f-^i^o)  and 


Figure  4.29  Illustration  of  the 
proof  of  Theorem  3.2.  The 
constraints  gi(x)  =  c\  and 
g2(x)  =  C2  are  the  surfaces  S\  and 
S2  •  Any  extremum  of  /  must  occur 
at  points  where  V/  is  in  the  plane 
spanned  by  Vgi  and  Vg2. 


to  the  individual  hypersurfaces  S\,  .  ..  ,  Sk  whose  intersection  is  S.  It  follows  (via 
a  little  more  linear  algebra)  that  there  must  be  scalars  ki, . . . ,  kk  such  that 


V/(xo)  =  hVgi(xo)  +  A2Vg2(x0)  + 


■kkWgk(x0). 


A  suggestion  of  the  geometry  of  this  proof  is  provided  by  Figure  4.29  (where 
k  =  2  and  n  =  3).  ■ 

EXAMPLE  5  Suppose  the  cone  z2  =  x2  +  y2  is  sliced  by  the  plane  z  =  x  + 
y  +  2  so  that  a  conic  section  C  is  created.  We  use  Lagrange  multipliers  to  find 
the  points  on  C  that  are  nearest  to  and  farthest  from  the  origin  in  R3 . 

The  problem  is  to  find  the  minimum  and  maximum  distances  from  (0,  0,  0) 
of  points  (x,  y,  z)  on  C.  For  algebraic  simplicity,  we  look  at  the  square  of  the 
distance  rather  than  the  actual  distance.  Thus,  we  desire  to  find  the  extrema  of 

f(x,  y,  Z)  =  x2  +  y2  +  z2 

(the  square  of  the  distance  from  the  origin  to  (x,  y,  z))  subject  to  the  constraints 

gi(x,  y,  z)  =  x2  +  y2  -  z2  =  0 
g2(x,  y,z)  =  x  +  y  -  z  =  -2 


Note  that 

Vgi(x,  y,  z)  =  (2x,  2y, 


-2z)    and    Vg2(x,y,z)  =  (1,1,-1). 


These  vectors  are  linearly  dependent  only  when  x  =  y  =  z-  However,  no  point  of 
the  form  (x,  x,  x)  simultaneously  satisfies  g\  =  0  and  g2  =  —2.  Hence,  Vgi  and 
Vg2  are  linearly  independent  at  all  points  that  satisfy  the  two  constraints.  There- 
fore, by  Theorem  3.2,  we  know  that  any  constrained  critical  points  (x0,  y0,  Zo) 
must  satisfy 

V/(*o,  yo,  z0)  =  kiVgi(x0,  y0,  zo)  +  k2Vg2(x0,  yo,  zo), 
as  well  as  the  two  constraint  equations.  Thus,  we  must  solve  the  system 

2x  =  2k\x  +  k2 
2y  =  2kiy  +  k2 
2z  =  -2k\z  —  k2  . 
x2  +  y2-z2  =  0 
x  +  y  -  z  =  -2 


Eliminating  k2  from  the  first  two  equations  yields 

k2  =  2x  —  2k\x  =2y  —  2A.iv, 


which  implies  that 


Therefore,  either 


2(x  -  y)(l  -  M)  =  0. 


x  =  y    or    A.1  =  1. 


The  condition  ki  =  1  implies  immediately  k2  =  0,  and  the  third  equation  of  the 
system  becomes  2z  =  —2z,  so  z  must  equal  0.  If  z  =  0,  then  x  and  y  must  be 


286       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 

zero  by  the  fourth  equation.  However,  (0,  0,  0)  is  not  a  point  on  the  plane  z  = 
x  +  y  +  2.  Thus,  the  condition  X \  =  1  leads  to  no  critical  point.  On  the  other  hand, 
if  x  =  y,  then  the  constraint  equations  (the  last  two  in  the  original  system  of  five) 
become 

2x2  -z2  =  0 
2x  —  z  =  —2 

Substituting  z  =  2x  +  2  yields 

2x2  -  (2x  +  2)2  =  0, 

equivalent  to 

2x2  +  8x  +  4  =  0, 

whose  solutions  are  x  =  —  2  ±  V2-  Therefore,  there  are  two  constrained  critical 
points 

ai  =  (-2  +  72,  -2  +  4l,  -2  +  2V2) 

and 

a2  =  (-2  -  V2,  -2  -  V2,  -2  -  2V2)  . 

We  can  check  that 

/(ai)  =  24  -  16 V2,        /(a2)  =  24  +  16^2, 

so  it  seems  that  ai  must  be  the  point  on  C  lying  nearest  the  origin,  and  a2  must  be 
the  point  that  lies  farthest.  However,  we  don't  know  a  priori  if  there  is  a  farthest 
point.  If  the  conic  section  C  is  a  hyperbola  or  a  parabola,  then  there  is  no  point 
that  is  farthest  from  the  origin.  To  understand  what  kind  of  curve  C  is,  note 
that  ai  has  positive  z-coordinate  and  a2  has  negative  z-coordinate.  Therefore,  the 
plane  z  =  x  +  y  +  2  intersects  both  nappes  of  the  cone  z2  =  x2  +  y2.  The  only 
conic  section  that  intersects  both  nappes  of  a  cone  is  a  hyperbola.  Hence,  C  is  a 
hyperbola,  and  we  see  that  the  point  ai  is  indeed  the  point  nearest  the  origin,  but 
the  point  a2  is  not  the  farthest  point.  Instead  a2  is  the  point  nearest  the  origin  on 
the  branch  of  the  hyperbola  not  containing  aj.  That  is,  local  constrained  minima 
occur  at  both  ai  and  a2,  but  only  &\  is  the  site  of  the  global  minimum.  (See 
Figure  4.30.)  ♦ 

A  Hessian  Criterion  for  Constrained  Extrema  (optional)  — 

As  Example  5  indicates,  it  is  often  possible  to  determine  the  nature  of  a  critical 
point  (constrained  or  unconstrained)  from  considerations  particular  to  the  prob- 
lem at  hand.  Sometimes  this  is  not  difficult  to  do  in  practice  and  can  provide 
useful  insight  into  the  problem.  Nonetheless,  occasionally  it  is  advantageous  to 
have  a  more  automatic  means  of  discerning  the  nature  of  a  constrained  critical 
point.  We  therefore  present  a  Hessian  criterion  for  constrained  critical  points. 
Like  the  one  in  the  unconstrained  case,  this  criterion  only  determines  the  local 
nature  of  a  critical  point.  It  does  not  provide  information  about  global  constrained 
extrema.2 


Figure  4.30  The 

point  ai  is  the  point 
on  the  hyperbola 
closest  to  the  origin. 
The  point  a2  is  the 
point  on  the  lower 
branch  of  the 
hyperbola  closest 
to  the  origin. 


2  We  invite  the  reader  to  consult  D.  Spring,  Amer.  Math.  Monthly,  92  (1985),  no.  9,  631-643  for  a  more 
complete  discussion. 


4.3  |  Lagrange  Multipliers  287 


In  general,  the  context  for  the  Hessian  criterion  is  this:  We  seek  extrema  of  a 
function  /:XcR"-^R  subject  to  the  k  constraints 


gi(xi,x2, . .  .,x„)  =  a 

g2(X\,X2,  X„)  =  C2 


gk(X\,X2, 


i  Xfi)  —  Ck 


We  assume  that  f,gi,...,gk  are  all  of  class  C2,  and  assume,  for  simplicity,  that 
/  and  the  gj 's  all  have  the  same  domain  X.  Finally,  we  assume  that  Vg\ ,  . . . ,  Vgk 
are  linearly  independent  at  the  constrained  critical  point  a.  Then,  by  Theorem  3.2, 
any  constrained  extremum  a  must  satisfy 

V/(a)  =  A.i  Vgi(a)  +  A2Vg2(a)  +  ■  ■  •  +  A.tVgt(a) 

for  some  scalars  Ai , . . . ,  A*.  We  can  consider  a  constrained  critical  point  to  be  a 
pair  of  vectors 

(A;a)  =  (Ai,  ...,Xk;au  ...,an) 

satisfying  the  aforementioned  equation.  In  fact,  we  can  check  that  (A;  a)  is  an 
unconstrained  critical  point  of  the  so-called  Lagrangian  function  L  defined 
by 


L(h,  ...,lk;xi,...,x„)=  f{x\,  . . .  ,xn)  -  y^Ji(gj(xi,  ...,xn)-  a). 


The  Hessian  criterion  comes  from  considering  the  Hessian  of  L  at  the  critical 
point  (A;  a).  Before  we  give  the  criterion,  we  note  the  following  fact  from  linear 
algebra:  Since  Vgi(a), . . . ,  Vgj(a)  are  assumed  to  be  linearly  independent,  the 
derivative  matrix  of  g  =  (g\ , . . . ,  gk)  at  a, 


Dg(a)  = 


dx\ 


(a) 


dx 


(a) 


9gi 
dxn 

dgk 
dxn 


(a) 


(a) 


hasa^:  x  k  submatrix(  obtained  by  deleting  n  —  k  columns  ofDg(a))  with  nonzero 
determinant.  By  relabeling  the  variables  if  necessary,  we  will  assume  that 


3jc 


(a) 


dxk 


(a) 


det 


^0 


dx 


(a) 


dxk 


(a) 


(i.e.,  that  we  may  delete  the  last  n  —  k  columns). 


Chapter  4  I  Maxi  ma  and  M 


inima  in  Several  Variables 


Second  derivative  test  for  constrained  local  extrema.  Given  a  constrained 
critical  point  a  of  /  subject  to  the  conditions  g\(x)  =  c\,  g2(x)  =  c2,  ■ . . , 
gk(x)  =  Ck,  consider  the  matrix 


HL(X;  a)  = 


0 


dx 


(a) 


dxn 


(a) 


0 

0 

dgk 
dx\ 

dgk 
dxn 


dx\ 


(a) 


dxn 


(a) 


(a) 


(a) 


dx\ 


(a) 


dx„ 


(a) 


where 


htj  = 


d2f 


(a)  -  h  -    -   (a)  -  X2  (a) 

OXj  dXi 


— —(a). 


dxjdxj  dxjdxi  dxjdxt  dxjdxt 

(Note  that  HL(k;  a)  is  an  («  +  /:)  x     +  /c)  matrix.)  By  relabeling  the  vari- 


ables as  necessary,  assume  that 


det 


3jc 


(a) 


3x 


(a) 


3*<t 


3xi. 


(a) 


(a) 


^0. 


As  in  the  unconstrained  case,  let  be  the  upper  leftmost  j  x  j  subma- 
trix  of  HL(X,  a).  For  j  =  1, 2, . . . ,  k  +  «,  let  <2;-  =  det  and  calculate  the 
following  sequence  of «  —  k  numbers: 

(-1)^+2,...,  (1) 

Note  that,  if  k  >  1,  the  sequence  in  (1)  is  not  the  complete  sequence  of  prin- 
cipal minors  of  HL(X,  a).  Assume  *4+n  =  det  HL(X,  a)  /  0.  The  numerical 
test  is  as  follows: 

1.  If  the  sequence  in  (1)  consists  entirely  of  positive  numbers,  then  /  has  a 
local  minimum  at  a  subject  to  the  constraints 


gi(x)  =  ci,     g2(x)  =  c2, 


gk(*)  =  ck. 


2.  If  the  sequence  in  (1)  begins  with  a  negative  number  and  thereafter  alter- 
nates in  sign,  then  /  has  a  local  maximum  at  a  subject  to  the  constraints 


gl(x)  =  Ci,      g2(x)  =  C2, 


gk(x)  =  Ck. 


3.  If  neither  case  1  nor  case  2  holds,  then  /  has  a  constrained  saddle  point 
at  a. 

In  the  event  that  det  HL(X,  a)  =  0,  the  constrained  critical  point  a  is  degen- 
erate, and  we  must  use  another  method  to  determine  whether  or  not  it  is  the 
site  of  an  extremum. 


4.3  |  Lagrange  Multipliers  289 


Finally,  in  the  case  of  no  constraint  equations  g,  (x)  =  c,  (i.e.,  k  =  0),  the 
preceding  criterion  becomes  the  usual  Hessian  test  for  a  function  /  of  n  variables. 

EXAMPLE  6   In  Example  1,  we  found  the  minimum  of  the  area  function 

A(x,  v,  z)  =  2xy  +  2yz  +  xz 
of  an  open  rectangular  box  subject  to  the  condition 

V(x,  y,  z)  =  xyz  =  4. 

Using  Lagrange  multipliers,  we  found  that  the  only  constrained  critical  point  was 
(2,  1 ,  2).  The  value  of  the  multiplier  A.  corresponding  to  this  point  is  2.  To  use  the 
Hessian  criterion  to  check  that  (2,  1,  2)  really  does  yield  a  local  minimum,  we 
construct  the  Lagrangian  function 

L(l;x,  y,  z)  =  A(x,  y,  z)  -  l(V(x,  y,  z)  -  4) 

=  2xy  +  2yz  +  xz  -  l{xyz  —  4). 


Then 


HL(l;x,y,z)  = 


0 

-yz 
-xz 
-xy 


-yz 
0 

2  —  Ix 
l-ly 


-xz 
2-lz 
0 

2  —  Ix 


-xy 
l-ly 
2  —  Ix 
0 


At  the  constrained  critical  point  (2;  2,  1 ,  2),  we  have 

HL{2;2,  1,2) 

The  sequence  of  determinants  to  consider  is 
(-l)1det//2(1)+1  = 


(-iydet//4 


0 

-2 

-4  " 

det 

-2 

0 

-2 

=  32, 

-4 

-2 

0 

-  0 

-2 

-4 

-2  " 

det 

-2 
-4 

0 

-2 

-2 
0 

-1 

-2 

_  -2 

-1 

-2 

0 

48. 


Since  these  numbers  are  both  positive,  we  see  that  (2,  1,2)  indeed  minimizes  the 
area  of  the  box  subject  to  the  constant  volume  constraint.  ♦ 

EXAMPLE  7  In  Example  5,  we  found  points  on  the  conic  section  C  denned 
by  equations 

\gl(x,y,z)  =  x2  +  y2-z2  =  0 
I  gi(x,  y,  z)  =  x  +  y  -  z  =  -2 
that  are  (constrained)  critical  points  of  the  "distance"  function 

f(x,y,z)  =  x2  +  y2  +  z2. 


To  apply  the  Hessian  criterion  in  this  case,  we  construct  the  Lagrangian  function 

z2)  -  m{x  +  y  -  z  +  2). 


L(l,m;x,  y,  z)  =  x2  +  y2  +  z2 


l(x2  +  y2 


290       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


The  critical  points  of  L,  found  by  setting  DL(l,  m;x,y,  z)  equal  to  0,  are 


(A.i;ai)  =  (-3  +  2V2, 

-24  + 

16^2 ; 

-2  +  wl, 

-2  +  wl,  - 

-2  + 

and 

(A2;a2)  =  (-3-2V2,- 

-24- 

16V2;- 

-2  -  y/2, 

-2-V2,  - 

-  2  - 

The  Hessian  of  L  is 

0 

0 

—2x 

-2y 

2z 

0 

0 

-1 

-1 

1 

HL(l,  m;x,  y,  z)  = 

— 2x 

-1 

2-2/ 

0 

0 

-2y 

-1 

0 

2-2/ 

0 

2z 

1 

0 

0  2 

+  21 

2V2). 


After  we  evaluate  this  matrix  at  each  of  the  critical  points,  we  need  to  compute 

(-l)2det//2(2)+1  =det//5. 

We  leave  it  to  you  to  check  that  for  (k\;  ai)  this  determinant  is  128  —  64 V2  f*s 
37.49,  and  for  (A2;  a2)  it  is  128  +  64V2  &  218.51.  Since  both  numbers  are  pos- 
itive, the  points  (—2  ±  V2>  —2  ±  -Jl,  —2  ±  2%/2)  are  both  sites  of  local  min- 
ima. By  comparing  the  values  of  /  at  these  two  points,  we  see  that  (—2  +  \/2, 
—2  +  V2,  —2  +  2*Jl)  must  be  the  global  minimum.  ♦ 


4.3  Exercises 


1 .  In  this  problem,  find  the  point  on  the  plane  2x  —  3 y  — 
z  =  4  that  is  closest  to  the  origin  in  two  ways: 

(a)  by  using  the  methods  in  §4.2  (i.e.,  by  finding  the 
minimum  value  of  an  appropriate  function  of  two 
variables); 

(b)  by  using  a  Lagrange  multiplier. 

In  Exercises  2-12,  use  Lagrange  multipliers  to  identify  the  crit- 
ical points  of  f  subject  to  the  given  constraints. 

2.  f(x,  y)  =  y,    2x2  +  y2  =  4 

3.  f(x,  y)  =  5x  +  2y,    5x2  +  2y2  =  14 

4.  f(x,  y)  =  xy,    2x  —  3y  =  6 

5.  f(x,  y,  z)  =  xyz,    2x  +  3 y  +  z  =  6 

6.  f(x,  y,z)  =  x2  +  y2  +  z2,    X  +  y  -  z  =  1 

7.  f(x,y,z)  =  3-x2-2y2-z2,    2x  +  y  +  z  =  2 

8.  f(x,  y,  z)  =  x6  +  y6  +  z6,    x2  +  y2  +  z2  =  6 

9.  f(x,  y,  z)  =  2x  +  y2-z2,    x  -  2y  =  0,  x  +  z  =  0 

10.  f(x,  y,  z)  =  2x  +  y2  +  2z,    x2  -  y2  =  1,  x  +  y  + 
z  =  2 

11.  f(x,  y,  z)  =  xy  +  yz,    x2  +  y2  =  1,  yz  =  1 


12.  f(x,  y,z)  =  x  +  y  +  z,    y2  -  x2  =  1,   x  +  2z=l 

13.  (a)  Find  the  critical  points  of  f(x,  y)  =  x2  +  y  sub- 

ject to  x2  +  2y2  =  1. 

(b)  Use  the  Hessian  criterion  to  determine  the  nature 
of  the  critical  point. 

14.  (a)  Find  any  critical  points  of  f(x,  y,  z,  w)  =  x2  + 

y2  +  z2  +  w2  subject  to  2x  +  y  +  z  =  1 ,  x  —  2z  — 
w  =  —2,  3x  +  y  +  2w  =  —  1. 

(b)  Use  the  Hessian  criterion  to  determine  the  nature 
of  the  critical  point.  (Note:  You  may  wish  to  use  a 
computer  algebra  system  for  the  calculations.) 

Just  as  sometimes  is  the  case  when  finding  ordinary  (i.e.,  un- 
constrained) critical  points  of  functions,  it  can  be  difficult  to 
solve  a  Lagrange  multiplier  problem  because  the  system  of 
equations  that  results  may  be  prohibitively  difficult  to  solve  by 
hand.  In  Exercises  15—19,  use  a  computer  algebra  system  to 
find  the  critical  points  of  the  given  function  f  subject  to  the 
constraints  indicated.  (Note:  You  may  find  it  helpful  to  provide 
numerical  approximations  in  some  cases.) 

^  15.  f(x,  y,  z)  =  3xy  -  4z,  3x  +  y  -  2xz  =  1 

^  16.  f(x,  y,  z)  =  3xy  -  4yz  +  5xz,     3x  +  y  +  2z=  12, 

2x  -  3y  +  5z  =  0 


4.3  I  Exercises  291 


17.  f(x,  y,  z)  =  y3  +  2xyz  -  x2,  x2  +  y2  +  z2  =  1 

18.  f(x,  y,  z)  =  x2  +  y2  -  xz2 ,  xy  +  z2  =  1 

19.  f(x,  y,  z,  w)  =  x2  +  y2  +  z2  +  w2,  x2  +  y2  =  1, 
x  +  y  +  z  +  w  =  \,x  —  y  +  z  —  w  =  0 

20.  Consider  the  problem  of  determining  the  extreme  val- 
ues of  the  function  f(x,  y)  =  x3  +  3y2  subject  to  the 
constraint  that  xy  =  —A. 

(a)  Use  a  Lagrange  multiplier  to  find  the  critical  points 
of  /  that  satisfy  the  constraint. 

(b)  Give  an  analytic  argument  to  determine  if  the  criti- 
cal points  you  found  in  part  (a)  yield  (constrained) 
maxima  or  minima  of  /. 

(c)  Use  a  computer  to  plot,  on  a  single  set  of  axes,  sev- 
eral level  curves  of  /  together  with  the  constraint 
curve  xy  =  —A.  Use  your  plot  to  give  a  geometric 
justification  for  your  answers  in  parts  (a)  and  (b). 

21.  Find  three  positive  numbers  whose  sum  is  18  and 
whose  product  is  as  large  as  possible. 

22.  Find  the  maximum  and  minimum  values  of 
f(x,  y,  z)  =  x  +  y  —  z  on  the  sphere  x2  +  y2  +  z2  = 
8 1 .  Explain  how  you  know  that  there  must  be  both  a 
maximum  and  a  minimum  attained. 

23.  Find  the  maximum  and  minimum  values  of  f(x,  y)  = 
x2  +  xy  +  v2  on  the  closed  disk  D  =  {(x,  y)  |  x2  +  y2 
<4}.  ' 

24.  You  are  sending  a  birthday  present  to  your  calculus  in- 
structor. Fly-By-Night  Delivery  Service  insists  that  any 
package  it  ships  be  such  that  the  sum  of  the  length  plus 
the  girth  be  at  most  1 08  in.  (The  girth  is  the  perimeter  of 
the  cross  section  perpendicular  to  the  length  axis — see 
Figure  4.31.)  What  are  the  dimensions  of  the  largest 
present  you  can  send? 


Figure  4.31  Diagram  for 
Exercise  24. 


25.  A  cylindrical  metal  can  is  to  be  manufactured  from 
a  fixed  amount  of  sheet  metal.  Use  the  method  of 
Lagrange  multipliers  to  determine  the  ratio  between 
the  dimensions  of  the  can  with  the  largest  capacity. 


26.  An  industrious  farmer  is  designing  a  silo  to  hold  her 
900tt  ft3  supply  of  grain.  The  silo  is  to  be  cylindrical 
in  shape  with  a  hemispherical  roof.  (See  Figure  4.32.) 
Suppose  that  it  costs  five  times  as  much  (per  square 
foot  of  sheet  metal  used)  to  fashion  the  roof  of  the  silo 
as  it  does  to  make  the  circular  floor  and  twice  as  much 
to  make  the  cylindrical  walls  as  the  floor.  If  you  were 
to  act  as  consultant  for  this  project,  what  dimensions 
would  you  recommend  so  that  the  total  cost  would  be 
a  minimum?  On  what  do  you  base  your  recommen- 
dation? (Assume  that  the  entire  silo  can  be  filled  with 
grain.) 


Figure  4.32  The  grain  silo 
of  Exercise  26. 

27.  You  are  in  charge  of  erecting  a  space  probe  on  the 
newly  discovered  planet  Nilrebo.  To  minimize  interfer- 
ence to  the  probe's  sensors,  you  must  place  the  probe 
where  the  magnetic  field  of  the  planet  is  weakest.  Nil- 
rebo is  perfectly  spherical  with  a  radius  of  3  (where 
the  units  are  thousands  of  miles).  Based  on  a  coordi- 
nate system  whose  origin  is  at  the  center  of  Nilrebo, 
the  strength  of  the  magnetic  field  in  space  is  given  by 
the  function  M(x,  y,  z)  =  xz  —  y2  +  3x  +  3.  Where 
should  you  locate  the  probe? 

28.  Heron's  formula  for  the  area  of  a  triangle  whose  sides 
have  lengths  x,  y,  and  z  is 

Area  =  ^s(s  —  x)(s  —  y)(s  —  z), 

where  s  =  j(x  +  y  +  z)  is  the  so-called  semi- 
perimeter  of  the  triangle.  Use  Heron's  formula  to  show 
that,  for  a  fixed  perimeter  P,  the  triangle  with  the 
largest  area  is  equilateral. 

29.  Use  a  Lagrange  multiplier  to  find  the  largest  sphere 
centered  at  the  origin  that  can  be  inscribed  in  the  ellip- 
soid 3x2  +  2y2  +  z2  =  6.  (Be  careful  with  this  prob- 
lem; drawing  a  picture  may  help.) 

30.  Find  the  point  closest  to  the  origin  and  on  the  line 
of  intersection  of  the  planes  2x  +  y  +  3z  =  9  and 
3x  +  2y  +  z  =  6. 


292       Chapter  4  I  Maxima  and  Minima  in  Several  Variables 


31 .  Find  the  point  closest  to  the  point  (2,5,-1)  and  on  the 
line  of  intersection  of  the  planes  x  —  2y  +  3z  =  8  and 

2z-y  =  3. 

32.  The  plane  x  +  y  +  z  =  4  intersects  the  paraboloid 
z  =  x2  +  y2  in  an  ellipse.  Find  the  points  on  the  el- 
lipse nearest  to  and  farthest  from  the  origin. 

33.  Find  the  highest  and  lowest  points  on  the  ellipse  ob- 


tained by  intersecting  the  paraboloid  z 
the  plane  x  +  y  +  1z  =  2, 


+  yz  with 


34.  Find  the  minimum  distance  between  a  point  on  the 
ellipse  x2  +  2y2  =  1  and  a  point  on  the  line  x  +  y  =  4. 
(Hint:  Consider  a  point  (x ,  y  )  on  the  ellipse  and  a  point 
(u,  v)  on  the  line.  Minimize  the  square  of  the  distance 
between  them  as  a  function  of  four  variables.  This  prob- 
lem is  difficult  to  solve  without  a  computer.) 

35.  (a)  Use  the  method  of  Lagrange  multipliers  to  find  crit- 

ical points  of  the  function  f(x,  y)  =  x  +  y  subject 
to  the  constraint  xy  =  6. 

(b)  Explain  geometrically  why  /  has  no  extrema  on 
the  set  {(x,  y)  \  xy  =  6}. 

36.  Leta,  fi,  and  y  denote  the  (interior)  angles  ofatriangle. 
Determine  the  maximum  value  of  sin  a  sin  /3  sin  y . 

37.  Let  S  be  a  surface  in  R3  given  by  the  equation 
g(x,  y,  z)  =  c,  where  g  is  a  function  of  class  C1  with 
nonvanishing  gradient  and  c  is  a  constant.  Suppose  that 
there  is  a  point  P  on  S  whose  distance  from  the  origin 
is  a  maximum.  Show  that  the  displacement  vector  from 
the  origin  to  P  must  be  perpendicular  to  S. 

38.  The  cylinder  x2  +  y2  =  4  and  the  plane  2x  +  2y  + 
z  =  2  intersect  in  an  ellipse.  Find  the  points  on 
the  ellipse  that  are  nearest  to  and  farthest  from  the 
origin. 

39.  Find  the  points  on  the  ellipse  3x2  —  Axy  +  3y2  =  50 
that  are  nearest  to  and  farthest  from  the  origin. 

40.  This  problem  concerns  the  determination  of  the  ex- 
trema of  f(x,  y)  =  ~Jx  +  subject  to  the  con- 
straint x2  +  y2  =  17,  where  x  >  0  and  y  >  0. 

(a)  Explain  why  /  must  attain  both  a  global  minimum 
and  a  global  maximum  on  the  given  constraint 
curve. 

(b)  Use  a  Lagrange  multiplier  to  solve  the  system  of 
equations 

\Vf(x,y)  =  XVg{x,y) 

U(x,  y)  =  o 

where  g(x,  y)  =  x2  +  y2.  You  should  identify  a 
single  critical  point  of  /. 

(c)  Identify  the  global  minimum  and  the  global  max- 
imum of  /  subject  to  the  constraint. 


41 .  Consider  the  problem  of  finding  extrema  of  f(x,  y)  = 
x  subject  to  the  constraint  y2  —  4x3  +  4x4  =  0. 

(a)  Use  a  Lagrange  multiplier  and  solve  the  system  of 
equations 

Wf(x,y)  =  Wg(x,y) 
[g(x,y)  =  0 

where  g(x,  y)  =  y2  —  4x3  +  4xA .  By  doing  so, 
you  will  identify  critical  points  of  /  subject  to  the 
given  constraint. 

(b)  Graph  the  curve  y2  —  4x3  +  4x4  =  0  and  use  the 
graph  to  determine  where  the  extrema  of  f(x,  y)  = 
x  occur. 

(c)  Compare  your  result  in  part  (a)  with  what  you 
found  in  part  (b).  What  accounts  for  any  differ- 
ences that  you  observed? 

42.  Consider  the  problem  of  finding  extrema  of 
f(x,  y,  z)  =  x2  +  y2  subject  to  the  constraint  z  =  c, 
where  c  is  any  constant. 

(a)  Use  the  method  of  Lagrange  multipliers  to  iden- 
tify the  critical  points  of  /  subject  to  the  constraint 
given  above. 

(b)  Using  the  usual  alphabetical  ordering  of  variables 
(i.e.,  x\  =x,x2  =  y,  x^  =  z),  construct  the  Hessian 
matrix  HL(X;a\,  02,  03)  (where  L(l;x,  y,  z)  = 
f{x,  y,  z)  —  l(z  —  c))  for  each  critical  point  you 
found  in  part  (a).  Try  to  use  the  second  deriva- 
tive test  for  constrained  extrema  to  determine  the 
nature  of  the  critical  points  you  found  in  part  (a). 
What  happens? 

(c)  Repeat  part  (b),  this  time  using  the  variable  order- 
ing x\  =  z,  X2  =  y,  X3  =  x.  What  does  the  second 
derivative  test  tell  you  now? 

(d)  Without  making  any  detailed  calculations,  discuss 
why  /  must  attain  its  minimum  value  at  the  point 
(0,  0,  c).  Then  try  to  reconcile  your  results  in  parts 
(b)  and  (c).  This  exercise  demonstrates  that  the  as- 
sumption that 


det 


dx 


dx 


(a) 


(a) 


dxk 


dxk 


(a) 


(a) 


#0 


is  important. 


43.  Consider  the  problem  of  finding  critical  points  of  the 
function  f(xi,  ■  ■  ■ ,  x„)  subject  to  the  set  of  k  con- 
straints 

gl(Xl,  ...,x„)  =  ci,     gi{xu  ...,*„)  =  c2, ... , 
gk(x\,  ...,x„)  =  ck. 
Assume  that  f,  gi,  g2,  •  ■  ■ ,  gk  are  all  of  class  C2. 


4.4  |  Some  Applications  of  Extrema 


(a)  Show  that  we  can  relate  the  method  of  Lagrange 
multipliers  for  determining  constrained  critical 
points  to  the  techniques  in  §4.2  for  finding  un- 
constrained critical  points  as  follows:  If 

(k,  a)  =  (ki,  . . . ,  kk;au  . . . ,  an) 

is  a  pair  consisting  of  k  values  for  Lagrange  mul- 
tipliers k\, . . . ,  kk  and  n  values  a\ , . . . ,  a„  for  the 
variables  X\, .. .  ,x„  such  that  a  is  a  constrained 
critical  point,  then  (k,  a)  is  an  ordinary  (i.e.,  un- 
constrained) critical  point  of  the  function 

LQi,  ...,lk;xi,  ...,x„) 

k 

=  f(X\,  ...,Xn)~  }     li(gi(Xl,  ...  ,Xn)-  Ci). 

1=1 

(b)  Calculate  the  Hessian  HL(k,  a),  and  verify  that  it 
is  the  matrix  used  in  §4.3  to  provide  the  criterion 
for  determining  the  nature  of  constrained  critical 
points. 


44.  The  unit  hypersphere  in  R"  (centered  at  the  origin 
0  =  (0, . . . ,  0))  is  defined  by  the  equation  xj+xj  + 

■  ■  ■  +  x\  =  1.  Find  the  pair  of  points  x  =  (xi  xn) 

and  y  =  (vi,  .  •  • ,  y„),  each  of  which  lies  on  the  unit 
hypersphere,  that  maximizes  and  minimizes  the  func- 
tion 

n 

f(x\,        Xn,  Vi,  .  .  •  ,  y„)  =  xtyt. 

i=l 

What  are  the  maximum  and  minimum  values  of  /? 

45.  Let  x  =  (xi, . . . ,  Xn)  and  y  =  (y\, . . . ,  yn)  be  any  vec- 
tors in  R"  and,  for =  1, . . . ,  n,  set 

u;  =  =    and    v,  =  =. 

(a)  Show  that  u  =  («i , . . . ,  un)  and  v  =  (i>i , . . . ,  vn) 
lie  on  the  unit  hypersphere  in  R" . 

(b)  Use  the  result  of  Exercise  44  to  establish  the 
Cauchy-Schwarz  inequality 

|x-y|  <  II x ||  ||y||. 


4.4  Some  Applications  of  Extrema 

In  this  section,  we  present  several  applications  of  the  methods  for  finding  both 
constrained  and  unconstrained  extrema  discussed  previously. 


53 
X 


Protein  level 


Figure  4.33  Height  versus  protein 
level. 


Protein  level 


Figure  4.34  Fitting  a  line  to 
the  data. 


Least  Squares  Approximation  

The  simplest  relation  between  two  quantities  x  and  y  is,  without  doubt,  a  linear 
one:  y  =  mx  +  b  (where  m  and  b  are  constants).  When  a  biologist,  chemist, 
psychologist,  or  economist  postulates  the  most  direct  connection  between  two 
types  of  observed  data,  that  connection  is  assumed  to  be  linear.  Suppose  that  Bob 
Biologist  and  Carol  Chemist  have  measured  certain  blood  protein  levels  in  an 
adult  population  and  have  graphed  these  levels  versus  the  heights  of  the  subjects 
as  in  Figure  4.33.  If  Prof.  Biologist  and  Dr.  Chemist  assume  a  linear  relationship 
between  the  protein  and  height,  then  they  desire  to  pass  a  line  through  the  data  as 
closely  as  possible,  as  suggested  by  Figure  4.34. 

To  make  this  standard  empirical  method  of  linear  regression  precise  (in- 
stead of  merely  graphical  and  intuitive),  we  first  need  some  notation.  Suppose  we 
have  collected  n  pairs  of  data  (x\ ,  yi),  (x2,  yi),  •  ■  ■ ,  (xn ,  y„ ).  (In  the  example  just 
described,  xt  is  the  protein  level  of  the  ith  subject  and  y,  his  or  her  height.)  We 
assume  that  there  is  some  underlying  relationship  of  the  form  y  =  mx  +  b,  and 
we  want  to  find  the  constants  m  and  b  so  that  the  line  fits  the  data  as  accurately 
as  possible.  Normally,  we  use  the  method  of  least  squares.  The  idea  is  to  find 
the  values  of  m  and  b  that  minimize  the  sum  of  the  squares  of  the  differences 
between  the  observed  y-values  and  those  predicted  by  the  linear  formula.  That  is, 
we  minimize  the  quantity 


D(m,  b)  =  [vi  -  {mx\  +  b)f  +  [y2  -  (mx2  +  b)f 
+  ---  +  [yn-(mxn+b)f, 


(1) 


294       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


Line  y  =  mx  +  b 


Figure  4.35  The  method  of  least  squares. 

where,  for  i  =  1, . . . ,  n,  y,-  represents  the  observed  y-value  of  the  data,  and 
mxi  +  b  represents  the  y-value  predicted  by  the  linear  relationship.  Hence,  each 
expression  in  D  of  the  form  y,  —  (mx,  +  b)  represents  the  error  between  the 
observed  and  predicted  y-values.  (See  Figure  4.35.)  They  are  squared  in  the 
expression  for  D  in  order  to  avoid  the  possibility  of  having  large  negative 
and  positive  terms  cancel  one  another,  thereby  leaving  little  or  no  "net  error," 
which  would  be  misleading.  Moreover,  D(m ,  b)  is  the  square  of  the  distance 
in  R"  between  the  point  (yi,  y2, . . . ,  y„)  and  the  point  {mx\  +  b,  mx2  +  b, . . . , 
mxn  +  b). 

Thus,  we  have  an  ordinary  minimization  problem  at  hand.  To  solve  it,  we 
need  to  find  the  critical  points  of  D.  First,  we  can  rewrite  D  as 

n 

D(m,  b)  =  y^[yj  -  (mxi  +  b)]2 

n  n  n  n 

=  ^2  yf  - 2m  X! Xtyi  ~2b^2, yi + XI  (mxi + fo)2- 
i=i       i=i  i=i 

H  ii 

1=1  1=1 

ii  n  n 

2  ^2  xi  Vi  +  2m      xf  +  2b  x\ 

i=l  i=l  1  =  1 

ii  n 

--  -2    yt  +  X! 2{,nXi  +  b"> 

1  =  1  i=l 

11  17 

=  —2       y,-  +  2m       X;  +  2nb. 

i  =  l  i=l 

When  we  set  both  partial  derivatives  equal  to  zero,  we  obtain  the  following  pair 
of  equations,  which  have  been  simplified  slightly: 

{H  x?) m  +  (H  xi)b  =  H  x>yi 

(2) 

(J2xj)m  +  nb  =  J2yi 

(All  sums  are  taken  from  i  =  1  to  «.)  Although  (2)  may  look  complicated,  it  is 
nothing  more  than  a  linear  system  of  two  equations  in  the  two  unknowns  m  and  b. 


Then 


i=i 


dD 

dm 


and 


3D 

~db 


4.4  |  Some  Applications  of  Extrema 


It  is  not  difficult  to  see  that  system  (2)  has  a  single  solution.  Therefore,  we  have 
shown  the  following: 


PROPOSITION  4.1  Given  n  data  points  (jci,  yi),  (x2,  yi), . . . ,  (x„,  y„)  with  not 
all  of  x\,  X2, . .  ■ ,  xn  equal,  the  function 

n 

D(m,b)  =  J2[yl-(mxl+b)f 
i=i 

has  a  single  critical  point  (mo,  bo)  given  by 

mo  =   j  ' 

and 

C>0  =   2  " 


Since  D(m ,  b)  is  a  quadratic  polynomial  in  m  and  b,  the  graph  of  z  =  D(m,b) 
is  a  quadric  surface.  (See  §2. 1 .)  The  only  such  surfaces  that  are  graphs  of  functions 
are  paraboloids  and  hyperbolic  paraboloids.  We  show  that,  in  the  present  case, 
the  graph  is  that  of  a  paraboloid  by  demonstrating  that  D  has  a  local  minimum  at 
the  critical  point  (mo,  &o)  given  in  Proposition  4.1. 

We  can  use  the  Hessian  criterion  to  check  that  D  has  a  local  minimum  at 
(mo,  bo).  We  have 


HD(m,  b)  = 


2  ^2  Xj  2n 


The  principal  minors  are  and  An  J2xf  ~  4(X>,)  •  The  first  minor  is 

obviously  positive,  but  determining  the  sign  of  the  second  requires  a  bit  more 
algebra.  (If  you  wish,  you  can  omit  reading  the  details  of  this  next  calculation 
and  rest  assured  that  the  story  has  a  happy  ending.)  Ignoring  the  factor  of  4,  we 

examine  the  expression  n     xf  —  (J2  *;)2-  Expanding  the  second  term  yields 


»£*?-(£■ 


= n  J2x>  -  [  Ex'2 + J22x'xj  ) 
i=i     \  i=i     i<j  ) 

n 

=  {n  -  l)^x'2  ~  ^22xixJ- 

i'=l  i<j 

On  the  other  hand  we  have 

n 

y^xxi  ~  xj^2  =  £  (xf  ~  2x'xj + x)) = (« - 1)  £  xf  -  £  2xjXj. 

(To  see  that  equation  (4)  holds,  you  need  to  convince  yourself  that 

£tf  +  *5)  =  0i-i)I>? 


(3) 


(4) 


296       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


by  counting  the  number  of  times  a  particular  term  of  the  form  x\  appears  in  the 
left-hand  sum.)  Thus,  we  have 

det HD(m,b)  =  4  I  n^x2  -  f^*, 


Best  fit 
line 


*  (5,  4) 


Figure  4.36  Data  for  the  linear 
regression  of  Example  1. 


=  4 1  (n  —  1)      x2  —       IxiXj  j    by  equation  (3), 
\  i=i         <</'  / 

=  4  /  ,(x,-  —  Xj)2    by  equation  (4). 


KJ 

Because  this  last  expression  is  a  sum  of  squares,  it  is  nonnegative.  Therefore,  the 
Hessian  criterion  shows  that  D  does  indeed  have  a  local  minimum  at  the  critical 
point.  Hence,  the  graph  of  z  =  D(m,  b)  is  that  of  a  paraboloid.  Since  the  (unique) 
local  minimum  of  a  paraboloid  is  in  fact  a  global  minimum  (consider  a  typical 
graph),  we  see  that  D  is  indeed  minimized  at  (m0,  b0). 

EXAMPLE  1  To  see  how  the  preceding  discussion  applies  to  a  specific  set  of 
data,  consider  the  situation  depicted  in  Figure  4.36. 

We  have  n  =  5,  and  the  function  D  to  be  minimized  is 

D(m,  b)  =  [2-  (m  +  b)f  +  [1  -  (2m  +  b)f  +  [5  -  (3m  +  b)]2 

+  [3  -  (4m  +  b)f  +  [4  -  (5m  +  b)f. 


We  compute 

=  15, 

Thus,  using  Proposition  4.1, 

5-51-15-15 


m 


5-55-15-15 


55, 

_  3 

"  5' 


15, 


55  ■  15  -  15  ■  51  _  6 
5-55-15-15  ~  5' 


The  best  fit  line  in  terms  of  least  squares  approximation  is 

3  6 
V  =  -x  H — . 
"^5  5 


Of  course,  linear  regression  is  not  always  an  appropriate  technique.  It  may 
not  be  reasonable  to  assume  that  the  data  points  fall  nearly  on  a  straight  line. 
Some  formula  other  than  y  =  mx  +  b  may  have  to  be  assumed  to  describe  the 
data  with  any  accuracy.  Such  a  postulated  relation  might  be  quadratic, 

y  =  ax2  +  bx  +  c, 

or  x  and  y  might  be  inversely  related 

a 

y=-+b. 

x 

You  can  still  apply  the  method  of  least  squares  to  construct  a  function  analogous 
to  D  in  equation  (1)  to  find  the  relation  of  a  given  form  that  best  fits  the  data. 

Another  way  that  least  squares  arise  is  if  y  depends  not  on  one  variable  but 
on  several:  x\,  X2,  ■  ■  ■ ,  xn.  For  example,  perhaps  adult  height  is  measured  against 


4.4  |  Some  Applications  of  Extrema 


Figure  4.37  A  particle  traveling 
in  a  force  field  F. 


blood  levels  of  10  different  proteins  instead  of  just  one.  Multiple  regression  is 
the  statistical  method  of  finding  the  linear  function 


y  =  a\X\  +  a2x2  + 
that  best  fits  a  data  set  of  (n  +  l)-tuples 


+  a„xn  +  b 


(a- 


(i) 


x1 


,xil\  yi),(*r.*2 


(2)  (2) 


{x 


(*)  VW 
1    '  x2  ' 


*,fU) 


We  can  find  such  a  "best  fit  hyperplane"  by  minimizing  the  sum  of  the  squares  of 
the  differences  between  the  y-values  furnished  by  the  data  set  and  those  predicted 
by  the  linear  formula.  We  leave  the  details  to  you.3 

Physical  Equilibria   


Let  F:  X  c  R3  — >  R3  be  a  continuous  force  field  acting  on  a  particle  that  moves 
along  a  path  x:  /  c  R  —>  R3  as  in  Figure  4.37.  Newton's  second  law  of  motion 
states  that 


F(x(f))  =  mx"(t), 


(5) 


where  m  is  the  mass  of  the  particle.  For  the  remainder  of  this  discussion,  we  will 
assume  that  F  is  a  gradient  field,  that  is,  that  F  =  —  VV  for  some  C1  potential 
function  V:X  c  R3  ->  R.  (See  §3.3  for  a  brief  comment  about  the  negative  sign.) 
We  first  establish  the  law  of  conservation  of  energy. 


THEOREM  4.2  (Conservation  of  energy)  Given  the  set-up  above,  the 
quantity 


\m\\x!{t)\\2  +  V(x(0) 


is  constant. 


The  term  hn  ||x'(?)  ||2  is  usually  referred  to  as  the  kinetic  energy  of  the  particle 
and  the  term  V(x(r))  as  the  potential  energy.  The  significance  of  Theorem  4.2 
is  that  it  states  that  the  sum  of  the  kinetic  and  potential  energies  of  a  particle 
is  always  fixed  (conserved)  when  the  particle  travels  along  a  path  in  a  gradient 
vector  field.  For  this  reason,  gradient  vector  fields  are  also  called  conservative 
vector  fields. 

Proof  of  Theorem  4.2  As  usual,  we  show  that  the  total  energy  is  constant  by 
showing  that  its  derivative  is  zero.  Thus,  using  the  product  rule  and  the  chain  rule, 
we  calculate 

d  r, 


dt  L 


i'(0  •  x'(0  +  V(x(/))]  =  mx"(t)  ■  x'(t)  +  V  V(x(0)  •  x'(0 

=  mx"(t)  •  x'(t)  -  F(x(0)  •  x'(t) 
=  mx"(t)  •  x'(t)  -  mx"(t)  •  x'(t) 
=  0, 

from  the  definitions  of  F  and  V  and  by  formula  (5). 


3  Or  you  might  consult  S.Weisberg,  Applied  Linear  Regression,  2nd  ed.,  Wiley-Interscience,  1985, 
Chapter  2.  Be  forewarned,  however,  that  to  treat  multiple  regression  with  any  elegance  requires  somewhat 
more  linear  algebra  than  we  have  presented. 


298       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


Figure  4.38  For  a  stable 
equilibrium  point,  the  path  of  a 
nearby  particle  with  a  sufficiently 
small  kinetic  energy  will  remain 
nearby  with  a  bounded  kinetic 
energy. 


In  physical  applications  it  is  important  to  identify  those  points  in  space  that 
are  "rest  positions"  for  particles  moving  under  the  influence  of  a  force  field.  These 
positions,  known  as  equilibrium  points,  are  such  that  the  force  field  does  not  act 
on  the  particle  so  as  to  move  it  from  that  position.  Equilibrium  points  are  of  two 
kinds:  stable  equilibria,  namely,  equilibrium  points  such  that  a  particle  perturbed 
slightly  from  these  positions  tends  to  remain  nearby  (for  example,  a  pendulum 
hanging  down  at  rest)  and  unstable  equilibria,  such  as  the  act  of  balancing  a  ball 
on  your  nose.  The  precise  definition  is  somewhat  technical. 


DEFINITION  4.3  Let  F:  X  c  R"  ->  R"  be  any  force  field.  Then  x0  e  X  is 
called  an  equilibrium  point  of  F  if  F(xo)  =  0.  An  equilibrium  point  xo  is 
said  to  be  stable  if,  for  every  r,  e  >  0,  we  can  find  other  numbers  ro,  eo  >  0 
such  that  if  we  place  a  particle  at  position  x  with  ||  x  —  x0 1|  <  r0  and  provide  it 
with  a  kinetic  energy  less  than  eo,  then  the  particle  will  always  remain  within 
distance  r  of  xq  with  kinetic  energy  less  than  e. 


In  other  words,  a  stable  equilibrium  point  xo  has  the  following  property:  You 
can  keep  a  particle  inside  a  specific  ball  centered  at  xo  with  a  small  kinetic  energy 
by  starting  the  particle  inside  some  other  (possibly  smaller)  ball  about  x0  and 
imparting  to  it  some  (possibly  smaller)  initial  kinetic  energy.  (See  Figure  4.38.) 

THEOREM  4.4    For  a  C1  potential  function  V  of  a  vector  field  F  =  -V  V, 

1.  The  critical  points  of  the  potential  function  are  precisely  the  equilibrium 
points  of  F. 

2.  If  x0  gives  a  strict  local  minimum  of  V,  then  x0  is  a  stable  equilibrium  point 
ofF. 


EXAMPLE  2  The  vector  field  F  =  (-6x  -  2y  -  2)i  +  {-2x  -  Ay  +  2)j  is 
conservative  and  has 

V(x,  y)  =  3x2  +  2xy  +  2x  +  2y2  -  2y  +  4 

as  a  potential  function  (meaning  that  F  =  —  V  V,  according  to  our  current  sign 
convention).  There  is  only  one  equilibrium  point,  namely,  (—  | ,  |) .  To  see  if  it  is 
stable,  we  look  at  the  Hessian  of  V : 


6  2 
2  4 


The  sequence  of  principal  minors  is  6,  20.  By  the  Hessian  criterion,  (— |,  |)  is  a 
strict  local  minimum  of  V  and,  by  Theorem  4.4,  it  must  be  a  stable  equilibrium 
point  of  F.  ♦ 

Proof  of  Theorem  4.4  The  proof  of  part  1  is  straightforward.  Since  F  =  — VV, 
we  see  that  F(x)  =  0  if  and  only  if  W(x)  =  0.  Thus,  equilibrium  points  of  F  are 
the  critical  points  of  V. 

To  prove  part  2,  let  xo  be  a  strict  local  minimum  of  V  and  x:/^R"aC' 
path  such  that  x(?0)  =  x0  for  some  to  e  /.  By  conservation  of  energy,  we  must 
have,  for  all  t  6  /,  that 

im||x'(0H2  +  V(x(/))  =  jm\\x'(t0)\\2  +  V(x(t0)). 


4.4  |  Some  Applications  of  Extrema 


Figure  4.39  On  the  surface 
S  =  {x  ]  g{\)  =  c],  the  component 
of  F  that  is  tangent  to  S  at  x  is 
denoted  by  $(x). 


To  show  that  xo  is  a  stable  equilibrium  point,  we  desire  to  show  that  we  can  bound 
the  distance  between  x(/)  and  xo  =  x(fo)  by  any  amount  r  and  the  kinetic  energy 
by  any  amount  e.  That  is,  we  want  to  show  we  can  achieve 

||x(O-x0||  <  r 

(i.e.,  x(?)  e  Br(x0)  in  the  notation  of  §2.2)  and 

im||x'(OH2  <  e. 

As  the  particle  moves  along  x  away  from  xo,  the  potential  energy  must  increase 
(since  xo  is  assumed  to  be  a  strict  local  minimum  of  potential  energy),  so  the 
kinetic  energy  must  decrease  by  the  same  amount.  For  the  particle  to  escape  from 
B,  (x0),  the  potential  energy  must  increase  by  a  certain  amount.  If  eo  is  chosen  to 
be  smaller  than  that  amount,  then  the  kinetic  energy  cannot  decrease  sufficiently 
(so  that  the  conservation  equation  holds)  without  becoming  negative.  This  being 
clearly  impossible,  the  particle  cannot  escape  from  Br(xo).  ■ 

Often  a  particle  is  not  only  acted  on  by  a  force  field  but  also  constrained  to  lie 
in  a  surface  in  space.  The  set-up  is  as  follows:  F  is  a  continuous  vector  field  on  R3 
acting  on  a  particle  that  lies  in  the  surface  S  =  {x  6  R3  |  g(x)  =  c},  where  g  is  a 
C1  function  such  that  Vg(x)  ^  0  for  all  x  in  S.  Most  of  the  comments  made  in  the 
unconstrained  case  still  hold  true,  provided  F  is  replaced  by  the  vector  component 
of  F  tangent  to  S.  Since,  at  x  e  S,  Vg(x)  is  normal  to  S,  this  tangential  component 
of  F  at  x  is 

<Kx)  =  F(x)  -  projVi?(x)F(x).  (6) 
(SeeFigure4.39.)Theninplaceofformula(5),wehave,forapathx:  /  cR->S, 

$(x(f))  =  mx"(t).  (7) 
We  can  now  state  a  "constrained  version"  of  Theorem  4.4. 


THEOREM  4.5    For  a  C1  potential  function  V  of  a  vector  field  F  =  -  V  V, 

1.  If  V|s  has  an  extremum  at  xo  e  5,  then  xo  is  an  equilibrium  point  in  S. 

2.  If  V  |  s  has  a  strict  local  minimum  at  x0  e  S,  then  x0  is  a  stable  equilibrium 
point. 


Sketch  of  proof  Forpart  1,  if  V\s  has  an  extremum  at  xo,  then,  by  Theorem  3.1, 
we  have,  for  some  scalar  X,  that 

VV(x0)  =  AVg(x0). 

Hence,  because  F  =  —  VV, 

F(x0)  =  -AVg(xo), 

implying  that  F  is  normal  to  5  at  xo.  Thus,  there  can  be  no  component  of  F  tangent 
to  S  at  xo  (i.e.,  4>(xo)  =  0).  Since  the  particle  is  constrained  to  lie  in  S,  we  see 
that  the  particle  is  in  equilibrium  in  S. 

The  proof  of  part  2  is  essentially  the  same  as  the  proof  of  part  2  of  Theorem 
4.4.  The  main  modification  is  that  the  conservation  of  energy  formula  in  Theorem 
4.2  must  be  established  anew,  as  its  derivation  rests  on  formula  (5),  which  has 
been  replaced  by  formula  (7).  Consequently,  using  the  product  and  chain  rules, 


Chapter  4  i  Maxima  and  Minima  in  Several  Variables 


x2  +  y2  +  (z  -  2r)2 


(0, 0, 3r) 

=  r2-i 

f      F ' 

'(0,0./-) 

Figure  4.40  On  the  sphere 
x2  +  y2  +  (z-  2r)2  =  r2,  the 
points  (0,  0,  r)  and  (0,  0,  3r) 
are  equilibrium  points  for  the 
gravitational  force  field 
F  =  —  mgk. 


we  check,  for  x:  / 
d 
dt 


m\\x'(t)\\2  +  V(x(r))]  =  jt[^mx'(t).x'(t)+  V(x(t))] 


=  mx"(t)  •  x'(t)  +  VV(x(/))  •  x'(0- 


Then,  using  formula  (6),  we  have 
d 


dt 


m \\x'(t)\\2  +  V(x(0)]  =  x'(t)  •  mx"(t)  -  F(x(f ))  •  x'(t) 

=  x'(t).<t>(x(t))-F(x(t)).x'(t) 

=  x'(t)  ■  [F(x(0)  -  projVg(x(r))F(x(0)] 

-F(x(/))-x'(f) 
=  -x'(0-projVs(x(r))F(x(0) 


after  cancellation.  Thus,  we  conclude  that 
d 
dt 

since  x'(t )  is  tangent  to  the  path  in  S  and,  hence,  tangent  to  S  itself  at  x(t),  while 
projvo(X(j))F(x(0)  is  parallel  to  Vg(x(f ))  and,  hence,  perpendicular  to  S  at  x(r).  ■ 


m||x'(OII2  +  V(x(0)]  =  0, 


EXAMPLE  3  Near  the  surface  of  the  earth,  the  gravitational  field  is  ap- 
proximately 

F  =  —mgk. 

(We're  assuming  that,  locally,  the  surface  of  the  earth  is  represented  by  the  plane 
z  =  0.)  Note  that  F  =  -V V,  where 

V(x,  y,  z)  =  mgz. 

Now  suppose  a  particle  of  mass  m  lies  on  a  small  sphere  with  equation 

h(x,  y,  z)  =  x2  +  y2  +  (z-  2rf  =  r2. 

We  can  find  constrained  equilibria  for  this  situation,  using  a  Lagrange  multiplier. 
The  gradient  equation  V  V  =  XV h,  along  with  the  constraint,  yields  the  system 

0  =  2Ax 
0  =  2ky 

mg  =  2X(z  —  2r) 
x2  +  y2  +  (z  -  2r)2  =  r2 

Because  m  and  g  are  nonzero,  X  cannot  be  zero.  The  first  two  equations  imply 
x  =  y  =  0.  Therefore,  the  last  equation  becomes 


which  implies 


(z-2r)2  =  r2, 


Z  =  r,  3r 


are  the  solutions.  Consequently,  the  positions  of  equilibrium  are  (0,  0,  r)  and 
(0,  0,  3r)  (corresponding  to  X  =  —mg/2r  and  +mg/2r,  respectively).  From  ge- 
ometric considerations,  we  see  V  is  strictly  minimized  at  S  at  (0,  0,  r)  and  maxi- 
mized at  (0,  0,  3r)  as  shown  in  Figure  4.40.  From  physical  considerations,  (0,  0,  r) 


4.4  |  Some  Applications  of  Extrema 


is  a  stable  equilibrium  and  (0,  0,  3r)  is  an  unstable  one.  (Try  balancing  a  marble 


Applications  to  Economics 

We  present  two  illustrations  of  how  Lagrange  multipliers  occur  in  problems  in- 
volving economic  models. 

EXAMPLE  4  The  usefulness  of  amounts  x\,  x%, . . . ,  xn  of  (respectively)  dif- 
ferent capital  goods  G\,Gi,  ...,Gn  can  sometimes  be  measured  by  a  function 
U(x\,  X2,  ■ . . ,  xn),  called  the  utility  of  these  goods.  Perhaps  the  goods  are  indi- 
vidual electronic  components  needed  in  the  manufacture  of  a  stereo  or  computer, 
or  perhaps  U  measures  an  individual  consumer's  utility  for  different  commodi- 
ties available  at  different  prices.  If  item  G,  costs  a,  per  unit  and  if  M  is  the  total 
amount  of  money  allocated  for  the  purchase  of  these  n  goods,  then  the  consumer 
or  the  company  needs  to  maximize  U(x\,  x%  x„)  subject  to 


This  is  a  standard  constrained  optimization  problem  that  can  readily  be  approached 
by  using  the  method  of  Lagrange  multipliers. 

For  instance,  suppose  you  have  a  job  ordering  stationery  supplies  for  an  office. 
The  office  needs  three  different  types  of  products  a,  b,  and  c,  which  you  will  order 
in  amounts  x,  y,  and  z,  respectively.  The  usefulness  of  these  products  to  the  smooth 
operation  of  the  office  turns  out  to  be  modeled  fairly  well  by  the  utility  function 
U(x,  y,  z)  =  xy  +  xyz.  If  product  a  costs  $3  per  unit,  product  b  $2  a  unit,  and 
product  c  $  1  a  unit  and  the  budget  allows  a  total  expenditure  of  not  more  than 
$899,  what  should  you  do?  The  answer  should  be  clear:  You  need  to  maximize 

U(x,  y,  z)  =  xy  +  xyz    subject  to    B(x,  y,  z)  =  3x  +  2y  +  z  =  899. 

The  Lagrange  multiplier  equation,  VC/(x,  y,  z)  =  XVB(x,  y,  z),  and  the  budget 
constraint  yield  the  system 


The  last  equality  implies  that  either  x  =  0  or  y  =  (z  +  l)/2.  We  can  reject  the 
first  possibility,  since  t/(0,  y,  z)  =  0  and  the  utility  U(x,  y,  z)  >  0  whenever 
x,  y,  and  z  are  all  positive.  Thus,  we  are  left  with  y  =  (z  +  l)/2.  This  in  turn 
implies  that  A.  =  (z  +  l)2/6.  Substituting  for  y  in  the  constraint  equation  shows 
that  x  =  (898  —  2z)/3,  so  that  equation  xy  =  X  becomes 


on  top  of  a  ball.) 


♦ 


a^xi  +  02X2  +  ■  ■  ■  +  a„xn  =  M. 


y  +  yz  =  3A 
x  +  xz  =  2k 
xy  =  X 

3x  +  2y  +  z  =  899 


Solving  for  X  in  the  first  three  equations  yields 


which  is  satisfied  by  either  z  =  —  1  (which  we  reject)  or  by  z  =  299.  The  only 
realistic  critical  point  for  this  problem  is  (100,  150,  299).  We  leave  it  to  you  to 
check  that  this  point  is  indeed  the  site  of  a  maximum  value  for  the  utility.  ♦ 


Chapter  4  i  Maxima  and  Minima  in  Several  Variables 


EXAMPLE  5  In  1928,  C.W.Cobb  and  P.  M.Douglas  developed  a  simple  model 
for  the  gross  output  Q  of  a  company  or  a  nation,  indicated  by  the  function 

Q(K,L)  =  AK"L1-", 

where  K  represents  the  capital  investment  (in  the  form  of  machinery  or  other 
equipment),  L  the  amount  of  labor  used,  and  A  and  a  positive  constants  with 
0  <  a  <  1 .  (The  function  Q  is  known  now  as  the  Cobb-Douglas  production 
function.)  If  you  are  president  of  a  company  or  nation,  you  naturally  wish  to 
maximize  output,  but  equipment  and  labor  cost  money  and  you  have  a  total 
amount  of  M  dollars  to  invest.  If  the  price  of  capital  is  p  dollars  per  unit  and 
the  cost  of  labor  (in  the  form  of  wages)  is  w  dollars  per  unit,  so  that  you  are 
constrained  by 

B(K,  L)  =  pK  +  wL  <  M, 

what  do  you  do? 

Again,  we  have  a  situation  ripe  for  the  use  of  Lagrange  multipliers.  Before 
we  consider  the  technical  formalities,  however,  we  consider  a  graphical  solution. 
Draw  the  level  curves  of  Q,  called  isoquants,  as  in  Figure  4.41.  Note  that  Q 
increases  as  we  move  away  from  the  origin  in  the  first  quadrant.  The  budget 
constraint  means  that  you  can  only  consider  values  of  K  and  L  that  lie  inside  or 
on  the  shaded  triangle.  It  is  clear  that  the  optimum  solution  occurs  at  the  point 
(K,  L)  where  the  level  curve  is  tangent  to  the  constraint  line  pK  +  wL  =  M. 

Here  is  the  analytical  solution:  From  the  equation  VQ(K,  L)  =  XVB(K,  L) 
plus  the  constraint,  we  obtain  the  system 

'AaK"-lLl-a  =  Xp 
A(l  -  d)Kalra  =  Xw. 
pK  +  wL  =  M 


Solving  for  p  and  w  in  the  first  two  equations  yields 


K 


Figure  4.41  A  family  of  isoquants.  The  optimum  value  of  Q(K,  L)  subject  to  the 
constraint  pK  +  wL  =  M  occurs  where  a  curve  of  the  form  Q  =  c  is  tangent  to  the 
constraint  line. 


4.4  I  Exercises  303 


Substitution  of  these  values  into  the  third  equation  gives 

Aa    „   i  „     A(l  -  a)    „    ,  „ 

—  K  L      +  —  -K"Ll-a  =  M. 

X  X 

Thus, 

X  =  —KaL1-", 
M 

and  the  only  critical  point  is 

(Ma  M(\-a) 
\  P  w 

From  this  geometric  discussion,  we  know  that  the  critical  point  must  yield  the 
maximum  output  Q. 

From  the  Lagrange  multiplier  equation,  at  the  optimum  values  for  L  and  K, 
we  have 

p dK      w  dL 

This  relation  says  that,  at  the  optimum  values,  the  marginal  change  in  output  per 
dollar's  worth  of  extra  capital  equals  the  marginal  change  per  dollar's  worth  of 
extra  labor.  In  other  words,  at  the  optimum  values,  exchanging  labor  for  capital 
(or  vice  versa)  won't  change  the  output.  This  is  by  no  means  the  case  away  from 
the  optimum  values. 

There  is  not  much  that  is  special  about  the  function  Q  chosen.  Most  of  our 
observations  remain  true  for  any  C2  function  Q  that  satisfies  the  conditions 

dQ  dQ  >Q        d2Q  d2Q 
dK'  dL  ~    '        dK2,  dL2  < 

If  you  consider  what  these  relations  mean  qualitatively  about  the  behavior  of  the 
output  function  with  respect  to  increases  in  capital  and  labor,  you  will  see  that 
they  are  entirely  reasonable  assumptions.4  ♦ 


4.4  Exercises 


1.  Find  the  line  that  best  fits  the  following  data:  (0,  2), 
(1,3),  (2,5),  (3,3),  (4,2),  (5,7),  (6,7). 

2.  Show  that  if  you  have  only  two  data  points  (xi ,  y\ )  and 
(xi,  yi),  then  the  best  fit  line  given  by  the  method  of 
least  squares  is,  in  fact,  the  line  through  (x\,  yi)  and 
(xz,  yi)- 

3.  Suppose  that  you  are  given  «  pairs  of  data  (xi,  y\), 
(X2,  yz),  ■  ■  ■ ,  (x„,  yn)  and  you  seek  to  fit  a  function  of 
the  form  y  =  a/x  +  b  to  these  data. 

(a)  Use  the  method  of  least  squares  as  outlined  in  this 
section  to  construct  a  function  D(a,  b)  that  gives 
the  sum  of  the  squares  of  the  distances  between 
observed  and  predicted  j-values  of  the  data. 


(b)  Show  that  the  "best  fit"  curve  of  the  form  y  = 
a/x  +  b  should  have 

n^yi/xt  -(E  !/*)(£*■) 

2 

»E  i/*?-(£i/*) 

and 

b  _  (E  i/-*,2) (E  yj)  -  (E  im) (E  y.M-) 
»EV*?-(Ei/*)2 

(All  sums  are  from  i  =  1  to  «.) 


4  For  more  about  the  history  and  derivation  of  the  Cobb— Douglas  function,  consult  R.  Geitz,  "The  Cobb- 
Douglas  production  function,"  UMAP  Module  No.  509,  Birkhauser,  1981. 


304       Chapter  4  |  Maxima  and  Minima  in  Several  Variables 


4.  Find  the  curve  of  the  form  y  =  a/x  +  b  that  best  fits 
the  following  data:  (1 ,  0),  (2,  - 1),  (|,  1),  and  (3,  - 1) . 
(See  Exercise  3.) 

5.  Suppose  that  you  have  n  pairs  of  data  (jci  ,  ji), 
(xi,  yi),  ■  ■  ■ ,  (xn,  yn)  and  you  desire  to  fit  a  quadratic 
function  of  the  form  y  =  ax2  +  bx  +  c  to  the  data. 
Show  that  the  "best  fit"  parabola  must  have  coefficients 
a,  b,  and  c  satisfying 

(E  xty  +  (E*,3)&  +  (E-*,2)c  =  Z*fyi 
(E -v,3)«  +  (E 4)b  +  (E *)c  =  E xm  . 
(E^2)«  +  (E^)^  +  «c  =  E.v/ 

(All  sums  are  from  i  =  1  to  n.) 

6.  (Note:  This  exercise  will  be  facilitated  by  the  use  of 
a  spreadsheet  or  computer  algebra  system.)  Egbert 
recorded  the  number  of  hours  he  slept  the  night  before 
a  major  exam  versus  the  score  he  earned,  as  shown  in 
the  table  below. 

(a)  Find  the  line  that  best  fits  these  data. 

(b)  Find  the  parabola  y  =  ax2  +  bx  +  c  that  best  fits 
these  data.  (See  Exercise  5.) 

(c)  Last  night  Egbert  slept  6.8  hr.  What  do  your  an- 
swers in  parts  (a)  and  (b)  predict  for  his  score  on 
the  calculus  final  he  takes  today? 


Hours  of  sleep 

Test  score 

8 

85 

8.5 

72 

9 

95 

7 

68 

4 

52 

8.5 

75 

7.5 

90 

6 

65 

7.  Let  F  =  {-2x  -  2y  -  l)i  +  (-2x  -  6y  -  2)j. 

(a)  Show  that  F  is  conservative  and  has  potential 
function 

V(x,  y)  =  x2  +  2xy  +  3y2  +x  +  2y 

(i.e.,F  =  -VV). 

(b)  What  are  the  equilibrium  points  of  F?  The  stable 
equilibria? 

8.  Suppose  a  particle  moves  in  a  vector  field  F  in  R2  with 
physical  potential 

V(x,  y)  =  2x2  -  Sxy  -  y2  +  I2x  -  8y  +  12. 

Find  all  equilibrium  points  of  F  and  indicate  which,  if 
any,  are  stable  equilibria. 


9.  Let  a  particle  move  in  the  vector  field  F  in  R3  whose 
physical  potential  is  given  by 

V(x,  y,  z)  =  3x2  +  2xy  +  z2  -  2yz  + 3x  + 5y  -  10. 

Determine  the  equilibria  of  F  and  identify  those  that 
are  stable. 

10.  Suppose  that  a  particle  of  mass  m  is  constrained  to 
move  on  the  ellipsoid  2x2  +  3y2  +  z2  =  1  subject  to 
both  a  gravitational  force  F  =  —mgk.,  as  well  as  to  an 
additional  potential  V(x,  y,  z)  =  2x. 

(a)  Find  any  equilibrium  points  for  this  situation. 

(b)  Are  there  any  stable  equilibria? 

1 1 .  The  Sukolux  Vacuum  Cleaner  Company  manufactures 
and  sells  three  types  of  vacuum  cleaners:  the  standard, 
executive,  and  deluxe  models.  The  annual  revenue  in 
dollars  as  a  function  of  the  numbers  x,  y,  and  z  (re- 
spectively) of  standard,  executive,  and  deluxe  models 
sold  is 

R{x,  y,  z)  =  xyz2  -  25,000*  -  25,000y  -  25,000z. 

The  manufacturing  plant  can  produce  200,000  total 
units  annually.  Assuming  that  everything  that  is  manu- 
factured is  sold,  how  should  production  be  distributed 
among  the  models  so  as  to  maximize  the  annual 
revenue? 

1 2.  Some  simple  electronic  devices  are  to  be  designed  to 
include  three  digital  component  modules,  types  1,  2, 
and  3,  which  are  to  be  kept  in  inventory  in  respective 
amounts  X\,  x%,  and  xj,.  Suppose  that  the  relative  im- 
portance of  these  components  to  the  various  devices  is 
modeled  by  the  utility  function 

U (x\ ,  X%,  X3)  =  X\X2  +  2x\X^  +  X\X2Xt,. 

You  are  authorized  to  purchase  $90  worth  of  these  parts 
to  make  prototype  devices.  If  type  1  costs  $1  per  com- 
ponent, type  2  $4  per  component,  and  type  3  $2  per 
component,  how  should  you  place  your  order? 

13.  A  farmer  has  determined  that  her  cornfield  will  yield 
corn  (in  bushels)  according  to  the  formula 

B(x,y)  =  4x2  +  y2  +  600, 

where  x  denotes  the  amount  of  water  (measured  in 
hundreds  of  gallons)  used  to  irrigate  the  field  and  y 
the  number  of  pounds  of  fertilizer  applied  to  the  field. 
The  fertilizer  costs  $10  per  pound  and  water  costs  $15 
per  hundred  gallons.  If  she  can  allot  $500  to  prepare 
her  field  through  irrigation  and  fertilization,  use  a 
Lagrange  multiplier  to  determine  how  much  water  and 
fertilizer  she  should  purchase  in  order  to  maximize  her 
yield. 

14.  A  textile  manufacturer  plans  to  produce  a  cashmere/ 
cotton  fabric  blend  for  use  in  making  sweaters.  The 
amount  of  fabric  that  can  be  produced  is  given  by 

f(x,y)  =  4xy-2x-8y  +  3, 


True/False  Exercises  for  Chapter  4 


where  x  denotes  the  number  of  pounds  of  raw  cash- 
mere used  is  and  y  is  the  number  of  pounds  of  raw 
cotton.  Cotton  costs  $2  per  pound  and  cashmere  costs 
$8  per  pound. 

(a)  If  the  manufacturer  can  spend  $  1 000  on  raw  mate- 
rials, use  a  Lagrange  multiplier  to  advise  him  how 
he  should  adjust  the  ratio  of  materials  in  order  to 
produce  the  most  cloth. 

(b)  Now  suppose  that  the  manufacturer  has  a  budget 
of  B  dollars.  What  should  the  ratio  of  cotton  to 
cashmere  be  (in  terms  of  5)?  What  is  the  limiting 
value  of  this  ratio  as  B  increases? 

15.  The  CEO  of  the  Wild  Widget  Company  has  decided 
to  invest  $360,000  in  his  Michigan  factory.  His  eco- 
nomic analysts  have  noted  that  the  output  of  this 
factory  is  modeled  by  the  function  Q(K,  L)  = 
60  K1^  L2/3 ,  where  K  represents  the  amount  (in  thou- 


sands of  dollars)  spent  on  capital  equipment  and  L 
represents  the  amount  (also  in  thousands  of  dollars) 
spent  on  labor. 

(a)  How  should  the  CEO  allocate  the  $360,000 
between  labor  and  equipment? 

(b)  Check  that  BQ/dK  =  dQ/dL  at  the  optimal 
values  for  K  and  L. 

16.  Let  Q{K,L)  be  a  production  function  for  a  com- 
pany where  K  and  L  represent  the  respective  amounts 
spent  on  capital  equipment  and  labor.  Let  p  denote  the 
price  of  capital  equipment  per  unit  and  w  the  cost  of 
labor  per  unit.  Show  that,  subject  to  a  fixed  produc- 
tion Q(K,  L)  =  c,  the  total  cost  M  of  production  is 
minimized  when  K  and  L  are  such  that 

j_3G  _  j_ae 

p  dK      w  dL 


True/False  Exercises  for  Chapter  4 


1 .  If  /  is  a  function  of  class  C2  and  p%  denotes  the  second- 
order  Taylor  polynomial  of  /  at  a,  then  /(x)  ~  P2W 
when  x  ~  a. 

2.  The  increment  A/  of  a  function  f(x,  y)  measures  the 
change  in  the  z-coordinate  of  the  tangent  plane  to  the 
graph  of  /. 

3.  The  differential  df  of  a  function  f(x,  y)  measures  the 
change  in  the  z-coordinate  of  the  tangent  plane  to  the 
graph  of  /. 

4.  The  second-order  Taylor  polynomial  of  f(x,  y,  z)  = 
x2  +  3xz  +  y2  at  (1,  —1,  2)  is  pi(x,  y,  z)  =  x2  +  3xz 

+  y2- 

5.  The  second-order  Taylor  polynomial  of  f(x,y)  = 
x3  +  2xy  +  y  at  (0,  0)  is  pi(x,  y)  =  2xy  +  y. 

6.  The  second-order  Taylor  polynomial  of  f(x,y)  = 
x3  +  2xy  +  y  at  (1,  —1)  is  p2(x,  y)  =  2xy  +  y. 

7.  Near  the  point  (1,3,5),  the  function  f(x,y,z)  = 
3x4  +  2y3  +  z2  is  most  sensitive  to  changes  in  z. 

8.  The  Hessian  matrix  Hf(x\ , . . . ,  xn)  of  /  has  the  prop- 
erty that  Hf(Xl  ,...,xn)T  =  Hf(xx x„), 

9.  If  V/(ai , . . . ,  an)  =  0,  then  /  has  a  local  extremum 
at  a  =  (ai,  . . . ,  a„). 

10.  If  /  is  differentiable  and  has  a  local  extremum  at 
a  =  (ai, . . . ,  an),  then  V/(a)  =  0. 

11.  The  set  {(x,  y,  z)  \  4  <  x2  +  y2  +  z2  <  9}  is  compact. 

12.  The  set  {(x,  y)  \  2x  —  3y  =  1}  is  compact. 


13.  Any  continuous  function  f(x,  y)  must  attain  a  global 
maximum  on  the  disk  {(x,  y)  \  x2  +  y2  <  1}. 

14.  Any  continuous  function  f(x,y,z)  must  attain  a 
global  maximum  on  the  ball  {(x,  y,  z)  |  (x  —  l)2  + 
(y+l)2  +  z2<4). 

15.  If  f(x,  y)  is  of  class  C2,  has  a  critical  point  at  (a,  b), 
and  fxx(a,  b)fyy(a,  b)  —  fxy(a,  b)2  <  0,  then  /  has  a 
saddle  point  at  (a,  b). 

16.  If  det  ///(a)  =  0,  then  /  has  a  saddle  point  at  a. 

17.  The  function  f(x,  y,  z)  =  xiy2z  —  x2(y  +  z)  has  a 
saddle  point  at  (1 ,  —  1 ,  2). 

18.  The  function  f(x,  y,  z)  =  x2  +  y2  +  z2  —  yz  has  a  lo- 
cal maximum  at  (0,  0,  0). 

19.  The  function  f(x,  y,  z)  =  xy3  —  x2z  +  z  has  a  degen- 
erate critical  point  at  (—  1 ,  0,  0). 

20.  The  function  F(xx  ,...,.«„)  =  2{xx  -  l)2  -  3(x2  -  2)2 

_l_  1_  (— iyi+1(n  +  l)(xn  —  rif  has  a  critical  point  at 

(l,2,...,n). 

21 .  The  function  F(xi,  x„)  =  2(xi-  l)2  -  3(x2  -  2)2 
+  ■  ■  ■  +  (—  l)"+1(rc  +  l)(xn  —  n)2  has  a  minimum  at 
(l,2,...,n). 

22.  All  local  extrema  of  a  function  of  more  than  one  vari- 
able occur  where  all  partial  derivatives  simultaneously 
vanish. 

23.  All  points  a  =  («i, . . . ,  ai)  where  the  function 
f(x\,  ...  ,x„)  has  an  extremum  subject  to  the  con- 
straint that  g{x\, ...  ,xn)  —  c,  are  solutions  to  the 


Chapter  4  !  Maxima  and  Minima  in  Several  Variables 


system  of  equations 

9/ 
3.vi 


9xi 


X  - 


24.  Any  solution  (Ai 
equations 

9*1 


9x„  9x„ 
g(Xi,  ...,x„)  =  c 

. . . ,  Xk,x\, . . . ,  x„)  to  the  system  of 


dx\  dx\ 


  =  Ai          +  •  •  •  +  Ajt  

dxn  dx„  dx„ 

gi(xu  ■■,  X„)  =  C\ 


gl(xi,  ...,Xn)  =  Ck 

yields  a  point  (x\, . . . ,  xn)  that  is  an  extreme  value 
of  /  subject  to  the  simultaneous  constraints  g\  = 

a,  ...,gk  =  ck. 

25.  To  find  the  critical  points  of  the  function  f(x,  y,  z,  w) 
subject  to  the  simultaneous  constraints  g{x,  y,  z,  w)  = 


c,  h(x,  y ,  z,  w)  =  d,  k(x,  y,z,w)  =  e  using  the  tech- 
nique of  Lagrange  multipliers,  one  will  have  to  solve 
a  system  of  four  equations  in  four  unknowns. 

26.  Suppose  that  f(x,  y,  z)  and  g(x,  y,  z)  are  of  class  C1 
and  that  (xo ,  yo ,  zo)  is  a  point  where  /  achieves  a  maxi- 
mum value  subject  to  the  constraint  that  g(x  ,  y ,  z)  =  c 
and  that  Vg(xo,  yo,  Zo)  is  nonzero.  Then  the  level  set 
of  /  that  contains  (xo,  yo,  zo)  must  be  tangent  to  the 
level  set  S  =  {(x,  y,  z)  \  g{x,  y,  z)  =  c}. 

27.  The  critical  points  of  f(x,  y,  z)  =  xy  +  2xz  +  2yz 

subject  to  the  constraint  that  xyz  =  4  are  the  same 

as  the  critical  points  of  the  function  F(x,  y)  = 

8  8 
xv  +  -  +  -. 

x  y 

28.  Given  data  points  (3,  1),  (4,  10),  (5,  8),  (6,  12),  to  find 
the  best  fit  line  by  regression,  we  find  the  minimum 
value  of  the  function  D(m,  b)  =  (3m  +  b  —  l)2  + 
(4m  +  b  -  10)2  +  (5m  +  b  -  8)2  +  (6m  +  b  -  12)2. 

29.  All  equilibrium  points  of  a  gradient  vector  field 
are  minimum  points  of  the  vector  field's  potential 
function. 

30.  Given  an  output  function  for  a  company,  the  marginal 
change  in  output  per  dollar  investment  in  capital  is  the 
same  as  the  marginal  change  in  the  output  per  dollar 
investment  in  labor. 


Miscellaneous  Exercises  for  Chapter  4 


1.  Let  V  =  jtr2h,  where  r  ~  ro  and  h  ~  ho.  What  re- 
lationship must  hold  between  ro  and  ho  for  V  to  be 
equally  sensitive  to  small  changes  in  r  and  hi 

2.  (a)  Find  the  unique  critical  point  of  the  function 

f(xux2,  ...,xn)  =  c-*?-*l— -< 

(b)  Use  the  Hessian  criterion  to  determine  the  nature 
of  this  critical  point. 

3.  The  Java  Joint  Gourmet  Coffee  House  sells  top-of- 
the-line  Arabian  Mocha  and  Hawaiian  Kona  beans. 
If  Mocha  beans  are  priced  at  x  dollars  per  pound 
and  Kona  beans  at  y  dollars  per  pound,  then  mar- 
ket research  has  shown  that  each  week  approximately 
80  —  lOOx  +  40y  pounds  of  Mocha  beans  will  be  sold 
and  20  +  60x  —  35y  pounds  of  Kona  beans  will  be 
sold.  The  wholesale  cost  to  the  Java  Joint  owners  is 
$2  per  pound  for  Mocha  beans  and  $4  per  pound  for 
Kona  beans.  How  should  the  owners  price  the  coffee 
beans  in  order  to  maximize  their  profits? 

4.  The  Crispy  Crunchy  Cereal  Company  produces  three 
brands,  X,  Y,  and  Z,  of  breakfast  cereal.  Each  month, 
x,  y,  and  z  (respectively)  1000-box  cases  of  brands  X, 


Y,  and  Z  are  sold  at  a  selling  price  (per  box)  of  each 
cereal  given  as  follows: 


Brand 

No.  cases  sold 

Selling  price  per  box 

X 

X 

4.00  -  0.02x 

Y 

y 

4.50-0.05y 

Z 

z 

5.00-O.lOz 

(a)  What  is  the  total  revenue  R  if  x  cases  of  brand  X,  y 
cases  of  brand  Y,  and  z  cases  of  brand  Z  are  sold? 

(b)  Suppose  that  during  the  month  of  November,  brand 
X  sells  for  $3.88  per  box,  brand  Y  for  $4.25,  and 
brand  Z  for  $4.60.  If  the  price  of  each  brand  is  in- 
creased by  $0. 10,  what  effect  will  this  have  on  the 
total  revenue? 

(c)  What  selling  prices  maximize  the  total  revenue? 

5.  Find  the  maximum  and  minimum  values  of  the 
function 

f(x,  y,  z)  =  x  —  V3  y 

on  the  sphere  x2  +  y2  +  z2  =  4  in  two  ways: 
(a)  by  using  a  Lagrange  multiplier; 


Miscellaneous  Exercises  for  Chapter  4 


the  rectangle  without  overlapping,  except  along  their 
edges.  (See  Figure  4.42.) 


(b)  by  substituting  spherical  coordinates  (thereby  de- 
scribing the  point  {x,y,  z)  on  the  sphere  as  x  = 
2  sin  (p  cos  9,  y  =  2  sirup  sin#,  z  =  2cos<p)  and 
then  finding  the  ordinary  (i.e.,  unconstrained)  ex- 
trema  of  f(x(cp,  9),  y(<p,  9),  z((p,  9)). 

6.  Suppose  that  the  temperature  in  a  space  is  given  by  the 
function 

T(x,  y,z)  =  200xyz2. 

Find  the  hottest  point(s)  on  the  unit  sphere  in  two  ways: 

(a)  by  using  Lagrange  multipliers; 

(b)  by  letting  x  =  sirup  cos  6*,  y  =  sirup  sin  9,  z  = 
cos  <p  and  maximizing  T  as  a  function  of  the  two 
independent  variables  <p  and  9 .  (Note:  It  will  help  if 
you  use  appropriate  trigonometric  identities  where 
possible.) 

7.  Consider  the  function  f(x,  y)  =  (y  —  2x2)(y  —  xz). 

(a)  Show  that  /  has  a  single  critical  point  at  the  origin. 

(b)  Show  that  this  critical  point  is  degenerate.  Hence, 
it  will  require  means  other  than  the  Hessian  crite- 
rion to  determine  the  nature  of  the  critical  point  as 
a  local  extremum. 

(c)  Show  that,  when  restricted  to  any  line  that  passes 
through  the  origin,  /  has  a  minimum  at  (0,  0). 
(That  is,  consider  the  function  F(x)  =  f(x,  mx), 
where  m  is  a  constant  and  the  function  G(y)  = 

f(0,  y).) 

(d)  However,  show  that,  when  restricted  to  the 
parabola  y  =  |jc2,  the  function  /  has  a  global 
maximum  at  (0,  0).  Thus,  the  origin  must  be  a  sad- 
dle point. 

(e)  Use  a  computer  to  graph  the  surface  z  =  f(x,  y). 

8.  (a)  Find  all  critical  points  of  f(x,  y)  =  xy  that  satisfy 

x2  +  y2  =  1. 

(b)  Draw  a  collection  of  level  curves  of  /  and,  on  the 
same  set  of  axes,  the  constraint  curve  x2  +  y2  =  1 , 
and  the  critical  points  you  found  in  part  (a). 

(c)  Use  the  plot  you  obtained  in  part  (b)  and  a  geomet- 
ric argument  to  determine  the  nature  of  the  critical 
points  found  in  part  (a). 

9.  (a)  Find  all  critical  points  of  f(x,y,z)  =  xy  that 

satisfy  x2  +  y2  +  z2  =  1 . 

(b)  Give  a  rough  sketch  of  a  collection  of  level  surfaces 
of  /  and,  on  the  same  set  of  axes,  the  constraint 
surface  x2  +  y2  +  z2  =  1 ,  and  the  critical  points 
you  found  in  part  (a). 

(c)  Use  part  (b)  and  a  geometric  argument  to  determine 
the  nature  of  the  critical  points  found  in  part  (a). 

10.  Find  the  area  A  of  the  largest  rectangle  so  that  two 
squares  of  total  area  1  can  be  placed  snugly  inside 


Figure  4.42  Figure  for  Exercise  10. 

1 1 .  Find  the  minimum  value  of 

f{x\  ,x2,  ... ,  x„)  =  x\  +  x\  H  h  x2 

subject  to  the  constraint  that  a\X\  +  a2x2  +  •  •  •  + 
a„xn  =  1,  assuming  that  a\  +  a\  +  •  •  •  +  a2  >  0. 

12.  Find  the  maximum  value  of 

f(xux2,  ...,xn)  =  (aiXi  +  a2x2  H  h  anxn)2, 

subject  to  x\  +  x\  +  •  •  •  +  x2  =  I.  Assume  that  not  all 
of  the  a,  's  are  zero. 

1 3.  Find  the  dimensions  of  the  largest  rectangular  box  that 
can  be  inscribed  in  the  ellipsoid  x2  +  2y2  +  4z2  =  12. 
Assume  that  the  faces  of  the  box  are  parallel  to  the 
coordinate  planes. 

14.  Your  company  must  design  a  storage  tank  for  Super 
Suds  liquid  laundry  detergent.  The  customer's  specifi- 
cations call  for  a  cylindrical  tank  with  hemispherical 
ends  (see  Figure  4.43),  and  the  tank  is  to  hold  8000  gal 
of  detergent.  Suppose  that  it  costs  twice  as  much  (per 
square  foot  of  sheet  metal  used)  to  machine  the  hemi- 
spherical ends  of  the  tank  as  it  does  to  make  the  cylin- 
drical part.  What  radius  and  height  do  you  recommend 
for  the  cylindrical  portion  so  as  to  minimize  the  total 
cost  of  manufacturing  the  tank? 


Figure  4.43  The  storage  tank 
of  Exercise  14. 

15.  Find  the  minimum  distance  from  the  origin  to  the 
surface  x2  —  (y  —  z)2  =  1 . 

1 6.  Determine  the  dimensions  of  the  largest  cone  that  can 
be  inscribed  in  a  sphere  of  radius  a. 

17.  Find  the  dimensions  of  the  largest  rectangular  box 
(whose  faces  are  parallel  to  the  coordinate  planes)  that 


Chapter  4  !  Maxima  and  Minima  in  Several  Variables 


can  be  inscribed  in  the  tetrahedron  having  three  faces  in 
the  coordinate  planes  and  fourth  face  in  the  plane  with 
equation  bcx  +  acy  +  abz  =  abc,  where  a,  b,  and  c 
are  positive  constants.  (See  Figure  4.44.) 


Figure  4.44  Figure  for  Exercise  17. 

1 8.  You  seek  to  mail  a  poster  to  your  friend  as  a  gift.  You 
roll  up  the  poster  and  put  it  in  a  cylindrical  tube  of  di- 
ameter a'  and  length  y.  The  postal  regulations  demand 
that  the  sum  of  the  length  of  the  tube  plus  its  girth  (i.e., 
the  circumference  of  the  tube)  be  at  most  108  in. 

(a)  Use  the  method  of  Lagrange  multipliers  to  find  the 
dimensions  of  the  largest- volume  tube  that  you  can 
mail. 

(b)  Use  techniques  from  single-variable  calculus  to 
solve  this  problem  in  another  way. 

1 9.  Find  the  distance  between  the  line  y  =  2x  +  2  and  the 
parabola  x  =  y2  by  minimizing  the  distance  between 
a  point  (x\ ,  yi)  on  the  line  and  a  point  (X2 ,  yi)  on  the 
parabola.  Draw  a  sketch  indicating  that  you  have  found 
the  minimum  value. 

20.  A  ray  of  light  travels  at  a  constant  speed  in  a  uniform 
medium,  but  in  different  media  (such  as  air  and  water) 
light  travels  at  different  speeds.  For  example,  if  a  ray  of 
light  passes  from  air  to  water,  it  is  bent  (or  refracted) 
as  shown  in  Figure  4.45.  Suppose  the  speed  of  light 

A 


\^  Medium 

1 

a  i  V 

i  V1! 

h 

i 

i 

Medium  2        10  £ 

i 

B 

in  medium  1  is  v\  and  in  medium  2  is  1)2 ■  Then,  by 
Fermat's  principle  of  least  time,  the  light  will  strike  the 
boundary  between  medium  1  and  medium  2  at  a  point 
P  so  that  the  total  time  the  light  travels  is  minimized. 

(a)  Determine  the  total  time  the  light  travels  in  going 
from  point  A  to  point  B  via  point  P  as  shown  in 
Figure  4.45. 

(b)  Use  the  method  of  Lagrange  multipliers  to  estab- 
lish Snell's  law  of  refraction:  that  the  total  travel 
time  is  minimized  when 


sin  6*i 
sin  6*2 


Vl 
!'2  ' 


(Hint:  The  horizontal  and  vertical  separations  of 
A  and  B  are  constant.) 

21.  Use  Lagrange  multipliers  to  establish  the  formula 
\ax0  +  by0  —  d\ 


D 


V«2  +  b2 


for  the  distance  D  from  the  point  (xq,  yo)  to  the  line 
ax  +  by  =  d. 

22.  Use  Lagrange  multipliers  to  establish  the  formula 

\ax0  +  by0  +  czo  -  d\ 


D 


si  a2  +  b2  +  c2 


for  the  distance  D  from  the  point  (xq,  yo,  zo)  to  the 
plane  ax  +  by  +  cz  =  d. 


23.  (a)  Show  that  the  maximum  value  of  f(x,y,z) 

2  2' 

xAyAz 


■2  subject  to  the  constraint  that  x2  +  y2  + 


2  2 
Z  =  a  is 


27     V  3 

(b)  Use  part  (a)  to  show  that,  for  all  x,  y,  and  z, 


(*Vz2)1/3 


x2  +  y2  +  z2 


(c)  Show  that,  for  any  positive  numbers  x\ ,  x% , . . . ,  xn , 

i  ^  xl+x2  +  ---+xn 

n 

The  quantity  on  the  right  of  the  inequality  is  the 
arithmetic  mean  of  the  numbers  x\,  X2,  ■  ■ . ,  xn, 
and  the  quantity  on  the  left  is  called  the  geomet- 
ric mean.  The  inequality  itself  is,  appropriately, 
called  the  arithmetic-geometric  inequality. 

(d)  Under  what  conditions  will  equality  hold  in  the 
arithmetic-geometric  inequality? 

In  Exercises  24-27  you  will  explore  how  some  ideas  from  ma- 
trix algebra  and  the  technique  of  Lagrange  multipliers  come 
together  to  treat  the  problem  of  finding  the  points  on  the  unit 
hypersphere 


Figure  4.45  Snell's  law  of  refraction. 


g(xu  ■■■,x„) 


+  x2  +  - 


l 


4.6  |  Miscellaneous  Exercises  for  Chapter  4 


that  give  extreme  values  of  the  quadratic  form 
f(*L 


where  the  Ojj  s  are  constants. 


24.  (a)  Use  a  Lagrange  multiplier  X  to  set  up  a  system  of 
n  +  1  equations  in  w  +  1  unknowns  xi , . . . ,  xn ,  X 
whose  solutions  provide  the  appropriate  con- 
strained critical  points. 

(b)  Recall  that  formula  (2)  in  §4.2  shows  that  the 
quadratic  form  /  may  be  written  in  terms  of  ma- 
trices as 


f(xi 


,x„) 


(1) 


where  the  vector  x  is  written  as  the  «  x  1  matrix 

X\ 

and  A  is  the  n  x  n  matrix  whose  ij th  entry 

is  atj .  Moreover,  as  noted  in  the  discussion  in  §4.2, 
the  matrix  A  may  be  taken  to  be  symmetric  (i.e.,  so 
that  AT  =  A),  and  we  will  therefore  assume  that 
A  is  symmetric.  Show  that  the  gradient  equation 
V/  =  XV g  is  equivalent  to  the  matrix  equation 


Ax  =  Xx. 


(2) 


Since  the  point  (xi, . . . ,  xn)  satisfies  the  constraint 
*]+•••+  x%  =  1,  the  vector  x  is  nonzero.  If  you 
have  studied  some  linear  algebra,  you  will  recog- 
nize that  you  have  shown  that  a  constrained  criti- 
cal point  (xi,  . . . ,  x„)  for  this  problem  corresponds 
precisely  to  an  eigenvector  of  the  matrix  A  asso- 
ciated with  the  eigenvalue  X. 


(c)  Now  suppose  that  x  : 


x\ 


is  one  of  the  eigen- 


vectors of  the  symmetric  matrix  A,  with  associated 
eigenvalue  X.  Use  equations  (1)  and  (2)  to  show,  if 
x  is  a  unit  vector,  that 


/(*! 


■  ,  Xn)  =  X. 


Hence,  the  (absolute)  minimum  value  that  / 
attains  on  the  unit  hypersphere  must  be  the  small- 
est eigenvalue  of  A  and  the  (absolute)  maximum 
value  must  be  the  largest  eigenvalue. 

25.  Let  n  =  2  in  the  situation  of  Exercise  24,  so  that  we  are 
considering  the  problem  of  finding  points  on  the  circle 


x2  +  y2  =  1  that  give  extreme  values  of  the  function 

f(x,  y)  =  ax2  +  2bxy  +  cy2 


[*  y] 


(a)  Find  the  eigenvalues  of  A 


a  b 
b  c 


by  identify- 


ing the  constrained  critical  points  of  the  optimiza- 
tion problem  described  above. 

(b)  Now  use  some  algebra  to  show  that  the  eigenval- 
ues you  found  in  part  (a)  must  be  real.  It  is  a  fact 
(that  you  need  not  demonstrate  here)  that  any  n  x  n 
symmetric  matrix  always  has  real  eigenvalues. 

26.  In  Exercise  25  you  noted  that  the  eigenvalues  X\,  X2 
that  you  obtained  are  both  real. 

(a)  Under  what  conditions  does  X\  =  X2I 

(b)  Suppose  that  X\  and  X2  are  both  positive.  Explain 
why  /  must  be  positive  on  all  points  of  the  unit 
circle. 

(c)  Suppose  that  Xi  and  X2  are  both  negative.  Explain 
why  /  must  be  negative  on  all  points  of  the  unit 
circle. 

27.  Let  /  be  a  general  quadratic  form  in  n  variables  de- 
termined by  an  n  x  n  symmetric  matrix  A,  that  is, 

f{x\, x„)  =       auxixj  =  xr^x- 

(a)  Show,  for  any  real  number  k,  that  f(kx\,..., 
kx„)  =  k2f(x\,  . . . ,  x„).  (This  means  that  a 
quadratic  form  is  a  homogeneous  polynomial  of 
degree  2 — see  Exercises  37-44  of  the  Miscella- 
neous Exercises  for  Chapter  2  for  more  about  ho- 
mogeneous functions.) 

(b)  Use  part  (a)  to  show  that  if  /  has  a  positive 
minimum  on  the  unit  hypersphere,  then  /  must 
be  positive  for  all  nonzero  x  e  R"  and  that  if  / 
has  a  negative  maximum  on  the  unit  hypersphere, 
then  /  must  be  negative  for  all  nonzero  x  €  R" . 
(Hint:  For  x  /  0,  let  u  =  x/||x||,  so  that  x  =  ku, 
where  k  =  ||x||.) 

(c)  Recall  from  §4.2  that  a  quadratic  form  /  is  said 
to  be  positive  definite  if  /(x)  >  0  for  all  nonzero 
x  e  R"  and  negative  definite  if  /(x)  <  0  for  all 
nonzero  x  e  R" .  Use  part  (b)  and  Exercise  24  to 
show  that  the  quadratic  form  /  is  positive  definite 
if  and  only  if  all  eigenvalues  of  A  are  positive,  and 
negative  definite  if  and  only  if  all  eigenvalues  of 
A  are  negative.  (Note:  As  remarked  in  part  (b)  of 
Exercise  25,  all  the  eigenvalues  of  A  will  be  real.) 


Multiple  Integration 


5.1  Introduction:  Areas  and 
Volumes 

5.2  Double  Integrals 

5.3  Changing  the  Order  of 
Integration 

5.4  Triple  Integrals 

5.5  Change  of  Variables 

5.6  Applications  of  Integration 

5.7  Numerical  Approximations 
of  Multiple  Integrals 
(optional) 

True/False  Exercises  for 
Chapter  5 

Miscellaneous  Exercises  for 
Chapter  5 

y 


Figure  5.1  The  graph  of 

y  =  f(x). 
y 


Figure  5.2  The  shaded  region  has 
area  /*  f(x)dx. 


5.1    Introduction:  Areas  and  Volumes 

Our  purpose  in  this  chapter  is  to  find  ways  to  generalize  the  notion  of  the  definite 
integral  of  a  function  of  a  single  variable  to  the  cases  of  functions  of  two  or 
three  variables.  We  also  explore  how  these  multiple  integrals  may  be  used  to 
meaningfully  represent  various  physical  quantities. 

Let  /  be  a  continuous  function  of  one  variable  defined  on  the  closed  interval 
[a,  b]  and  suppose  that  /  has  only  nonnegative  values.  Then  the  graph  of  /  looks 
like  Figure  5.1.  That  /  is  continuous  is  reflected  in  the  fact  that  the  graph  consists 
of  an  unbroken  curve.  That  /  is  nonnegative-valued  means  that  this  curve  does  not 
dip  below  the  x -axis.  We  know  from  one- variable  calculus  that  the  definite  integral 
fb  f(x)  dx  exists  and  gives  the  area  under  the  curve,  as  shown  in  Figure  5.2. 

Now  suppose  that  /  is  a  continuous,  nonnegative-valued  function  of  two 
variables  defined  on  the  closed  rectangle 

R  =  [(x,  y)  e  R2  |  a  <  x  <  b,  c  <  y  <  d] 

in  R2.  Then  the  graph  of  /  over  R  looks  like  an  unbroken  surface  that  never  dips 
below  the  xy-plane,  as  shown  in  Figure  5.3.  In  analogy  with  the  single-variable 
case,  there  should  be  some  sort  of  integral  that  represents  the  volume  under  the 
part  of  the  graph  that  lies  over  R.  (See  Figure  5.4.)  We  can  find  such  an  integral 
by  using  Cavalieri's  principle,  which  is  nothing  more  than  a  fancy  term  for  the 
method  of  slicing.  Suppose  we  slice  by  the  vertical  plane  x  =  xo,  where  xo  is  a 
constant  between  a  and  b.  Let  A(xo)  denote  the  cross-sectional  area  of  such  a 
slice.  Then,  roughly,  one  can  think  of  the  quantity  A(xo)  dx  as  giving  the  volume 
of  an  "infinitely  thin"  slab  of  thickness  dx  and  cross-sectional  area  A(xo).  (See 
Figure  5.5.)  Hence,  the  definite  integral 


V 


-f 

J  Li 


A(x)  dx 


gives  a  "sum"  of  the  volumes  of  such  slabs  and  can  be  considered  to  provide  a 
reasonable  definition  of  the  total  volume  of  the  solid. 

But  what  about  the  value  of  A(xo)?  Note  that  A(xo)  is  nothing  more  than  the 
area  under  the  curve  z  =  f(xo,  y),  obtained  by  slicing  the  surface  z  =  f(x,  y) 
with  the  plane  x  =  xq.  Therefore, 


A(x0) 


f(x0,  y)dy 


5.1  |  Introduction:  Areas  and  Volumes  311 


Figure  5.3  The  graph  of  Figure  5.4  The  region  under  the  Figure  5.5  A  slab  of  "volume" 

z  =  f(x,  v).  portion  of  the  graph  of  /  lying  dV  =  A(x0)  dx. 

over  R  has  volume  that  is  given 

by  an  integral. 


Plane  z  =  c 


(a,  0, 0) 


(0,6,0) 

y 

(a,  b,  0) 


Figure  5.6  Calculating  the 
volume  of  the  box  of  Example  1 . 


Figure  5.7  The  graph  of 
z  =  4  -  x2 
Example  2. 


4  -  x2  -  y2  of 


(remember  xq  is  a  constant),  and  so  we  find  that 


o  pupa 

=  /    A(x)(ix  =  /      /    /(x,  y)dy 

J  a  J  a     lJ  c 


dx. 


(i) 


The  right-hand  side  of  formula  (1)  is  called  an  iterated  integral.  To  calculate 
it,  first  find  an  "antiderivative"  of  f(x,  y)  with  respect  to  y  (by  treating  i  as  a 
constant),  evaluate  at  the  integration  limits  y  =  c  and  y  =  d,  and  then  repeat  the 
process  with  respect  to  x. 

EXAMPLE  1  Let's  make  sure  that  the  iterated  integral  defined  in  formula  (1) 
gives  the  correct  answer  in  a  case  we  know  well,  namely,  the  case  of  a  box.  We'll 
picture  the  box  as  in  Figure  5.6.  That  is,  the  box  is  bounded  on  top  and  bottom  by 
the  planes  z  =  c  (where  c  >  0)  and  z  =  0,  on  left  and  right  by  the  planes  y  =  0 
and  y  =  b  (where  b  >  0),  and  on  back  and  front  by  the  planes  x  =  0  and  x  =  a 
(a  >  0).  Hence,  the  volume  of  the  box  may  be  found  by  computing  the  volume 
under  the  graph  of  z  =  c  over  the  rectangle 

R  =  {(x,y)  |  0  <  x  <  a,0  <  y  <  b}. 

Using  formula  (1),  we  obtain 

V  =  y  j  cdydx  =  j  (^cy\-y~^  dx  =  j  cbdx  =  cbx\]zCQ  =  cba. 

This  result  checks  with  what  we  already  know  the  volume  to  be,  as  it  should.  ♦ 

EXAMPLE  2   We  calculate  the  volume  under  the  graph  of  z  =  4  —  x2  —  y2 

(Figure  5.7)  over  the  square 

R  =  {(x,  y)  |  -1  <  x  <  1,  -1  <  y  <  1}. 

Using  formula  (1)  once  again,  we  calculate  the  volume  by  first  integrating  with 
respect  to  y  (i.e.,  by  treating  x  as  a  constant  in  the  inside  integral)  and  then  by 


Chapter  5  |  Multiple  Integration 


y  =  y0  plane 


gure  5.8 

first. 


Y/y 

Slicing  by  y  =  yo 


integrating  w 


■/. 


th  respect  to  x.  The  details  are  as  follows: 
j  (4-x2-  y2)dydx  =  j 


dx 


y=-l 


A-xz 

8  -  2x2 
22       2  : 

 X  X 

3  3 


-4  +  x2  + 


dx 


dx 


-l 


22 
T 


22  2 
T  +  3 


40 

y 


In  our  development  of  formula  (1 ),  we  could  just  as  well  have  begun  by  slicing 
the  solid  with  the  plane  y  =  yo  (instead  of  with  the  plane  x  =  xq),  as  shown  in 
Figure  5.8.  Then,  in  place  of  formula  (1),  the  formula  that  results  is 


V 


rd  rb 

=  I    I    f(x,  y  )dxdy. 

J  c    J  a 


(2) 


Since  the  iterated  integrals  in  formulas  ( 1 )  and  (2)  both  represent  the  volume  of  the 
same  geometric  object,  we  can  summarize  the  preceding  discussion  as  follows. 


PROPOSITION  1 .1  Let  R  be  the  rectangle  {(x,  y)  \  a  <  x  <  b,  c  <  y  <  d)  and 
let  /  be  continuous  and  nonnegative  on  R.  Then  the  volume  V  under  the  graph 
of  /'  over  R  is 

pb    pel  pel  pb 

/    /    f(x,y)dydx=        /  f(x,y)dxdy. 

J a    J c  J  c     J  a 


EXAMPLE  3  We  find  the  volume  under  the  graph  of  z  =  cosx  siny  over  the 
rectangle 


R 


f  7T  TV  1 

=  \(x,y)  \  0<x  <-,  0<y  <-\ 


(See  Figure  5.9.)  From  formula  (1),  we  calculate  that  the  volume  is 


V 


fJt/2  fx/2 

=  /      cos  x  sin  y  dy  dx  =  / 

Jo     Jo  Jo 


*'2  (  sj2_ 
2 


(—  cosx  cos  j)|y=o/4  dx 


2-V2  r12 


— 2~ —  J 


2-  s/2 


sinx 


7T/2 


2  -  V2  2  -  s/2 

—^(1-0)  =  —^-. 


5.1  |  Exercises  313 


Figure  5.9  The  surface  z  =  cosx  siny  of 
Example  3. 


If  we  use  formula  (2)  instead  of  formula  (1),  we  obtain 

tjt/4    fji/2  rir/4 


Jo  Jo 


I  cos  x  sin  y  dx  dy  =  /  (sinx  sin  y)\xx=^2  dy 
Jo  Jo 

*n/4 


=  f 

Jo 


,t/4 


(sin  y  —  0)  dy  =  —  cos  y  |0 

4l  2-yfl 

=  ~-  (-1)  =  —  • 

2      V     '  2 

That  this  result  agrees  with  our  first  calculation  is  no  surprise  given 
Proposition  1.1.  ♦ 


5.1  Exercises 


Evaluate  the  iterated  integrals  given  in  Exercises  1-6. 


2.  J    J    y  sin  x  dy  dx 

3.  /     /    xey  dy  dx 

J -2  Jo 

rn/2  ri 

4.  /       /    ex  cos  y  dx  dy 
Jo  Jo 

5.  J  j  (ex+y  +  x2  +  In  y)  dx  dy 

6.  f    f  — ^—  dx  dy 

Jl  Ji  xy 


7.  Find  the  volume  of  the  region  that  lies  under  the  graph 
of  the  paraboloid  z  =  x2  +  y2  +2  and  over  the  rect- 
angle R  =  {(x,  y)  |  —  1  <  x  <  2,  0  <  y  <  2}  in  two 
ways: 

(a)  by  using  Cavalieri's  principle  to  write  the  volume 
as  an  iterated  integral  that  results  from  slicing  the 
region  by  parallel  planes  of  the  form  x  =  constant; 

(b)  by  using  Cavalieri's  principle  to  write  the  volume 
as  an  iterated  integral  that  results  from  slicing  the 
region  by  parallel  planes  of  the  form  y  =  constant. 

8.  Find  the  volume  of  the  region  bounded  on  top  by  the 
plane  z  =  x  +  3y  +  1,  on  the  bottom  by  the  xy-plane, 
and  on  the  sides  by  the  planes  x  =  0,  x  =  3,  y  =  1, 

y  =  2. 

9.  Find  the  volume  of  the  region  bounded  by  the  graph 
of  f{x,  y)  =  2x2  +  y4  sinjrx,  the  xy-plane,  and  the 
planes  x  =  0,  x  =  1,  y  =  —  1,  y  =  2. 


314       Chapter  5  |  Multiple  Integration 


In  Exercises  10—15,  calculate  the  given  iterated  integrals  and 
indicate  of  what  regions  in  R3  they  may  be  considered  to 
represent  the  volumes. 

10.  j  j  Idxdy 

11.  J  j  (16-x2  -y2)dydx 

/jr/2  />jt 
I    sin  x  cos  y  dx  dy 
-jr/2  JO 


13.  j  j  (4-x2)dxdy 


14.  J    J    \x\  sin  jry  dy  dx 

15.  J  J  (5-\y\)dxdy 

16.  Suppose  that  /  is  a  nonnegative- valued,  continu- 
ous function  defined  on  R  =  {(x,  y)  |  a  <  x  <  b,  c  < 
y  <  d}.  If  fix,  y)  <  M  for  some  positive  number  M, 
explain  why  the  volume  V  under  the  graph  of  /  over 
R  is  at  most  M(b  —  a)(d  —  c). 


5.2   Double  Integrals 

In  the  previous  section  we  saw  how  to  calculate  volumes  of  certain  solids  us- 
ing iterated  integrals.  The  ideas  were  mostly  straightforward,  but  the  situation 
we  addressed  was  rather  special:  We  only  solved  the  problem  of  computing  the 
volume  of  a  solid  defined  as  the  region  lying  under  the  graph  of  a  continuous, 
nonnegative-valued  function  f(x,  y)  and  above  a  rectangle  in  the  xv-plane.  It  is 
not  immediately  apparent  how  we  might  compute  the  volume  of  a  more  general 
solid  based  on  this  work. 

Thus,  in  this  section  we  define  a  more  general  notion  of  an  integral  of  a 
function  of  two  variables  that  will  allow  us  to  describe 

1.  integrals  of  arbitrary  functions  (i.e.,  functions  that  are  not  necessarily  non- 
negative  or  continuous)  and 

2.  integrals  over  arbitrary  regions  in  the  plane  (i.e.,  rather  than  integrals  over 
rectangles  only). 

We  focus  first  on  case  1 .  To  do  this,  we  start  fresh  with  some  careful  definitions 
and  notation.  The  ideas  involved  in  Definitions  2.1-2.3  below  are  different  from 
those  in  the  previous  section.  However,  we  will  see  that  there  is  a  key  connection 
(called  Fubini's  theorem)  between  the  notion  of  an  iterated  integral  discussed 
in  §5.1  and  that  of  a  double  integral,  which  will  be  described  in  Definition  2.3. 

The  Integral  over  a  Rectangle   

We  also  denote  a  (closed)  rectangle 

R  =  {{x,  y)  e  R2  |  a  <  x  <  b,  c  <  y  <  d} 

by  [a,  b]  x  [c,  d].  This  notation  is  intended  to  be  analogous  to  the  notation  for  a 
closed  interval. 


DEFINITION  2.1  Given  a  closed  rectangle  R  =  [a,  b]  x  [c,  d],  a  partition 
of  R  of  order  n  consists  of  two  collections  of  partition  points  that  break  up 
R  into  a  union  of  n2  subrectangles.  More  specifically,  for  i,  j  =  0,  . . . ,  n,  we 
introduce  the  collections  {*,  }  and  {yj},  so  that 

a  =  xq  <  x\  <  ■  ■  ■  <  x,_i  <  Xj  <  ■  ■  ■  <  x„  =  b, 


5.2  |  Double  Integrals  315 


and 


c  =  yo  <  yi  < 


< 


yj-y  <  yj 


<  yn  =  d. 


Let  Ax,-  =  x,-  —  Xi-i  (for  i  =  1, . . .  ,  n)  and  Ay;-  =  y,-  —  y;_i  (for  j  = 
1, . . . ,  n).  Note  that  Ax,-  and  Ayv-  are  just  the  width  and  height  (respec- 
tively) of  the  z'y'th  subrectangle  (reading  left  to  right  and  bottom  to  top)  of  the 
partition. 


An  example  of  a  partitioned  rectangle  is  shown  in  Figure  5.10.  We  do  not 
assume  that  the  partition  is  regular  (i.e.,  that  all  the  subrectangles  have  the  same 
dimensions). 


d  =  y„ 

yj-i 
c=y0 


+ 


H  h 


+ 


+ 


+ 


CI  —  Xq    X-[    X2  '  ■  ■  Xj  _  1     Xj   •  •  •  Xn  —  D 

Figure  5.10  A  partition  of  the  rectangle  [a,  b]  x  [c,  d]. 


DEFINITION  2.2  Suppose  that  /  is  any  function  denned  on  R  =  [a,  b]  x 
[c,  d]  and  partition  R  in  some  way.  Let  c,;-  be  any  point  in  the  subrectangle 

Rij  =  [xi-\,Xi\  x  [yj-uyj]    (i,j  =  l,...,  n). 

Then  the  quantity 

n 

where  AAjj  =  Ax,  Ay7  is  the  area  of  Rij,  is  called  a  Riemann  sum  of  /  on 
R  corresponding  to  the  partition. 


The  Riemann  sum 

S  =  £/(cy)AA(7 

ij 

depends  on  the  function  /,  the  choice  of  partition,  and  the  choice  of  the  "test 
point"  cy  in  each  subrectangle  Rij  of  the  partition.  The  Riemann  sum  itself  is 
just  a  weighted  sum  of  areas  AAjj  of  subrectangles  of  the  original  rectangle  R, 
the  weighting  being  given  by  the  value  /(cy). 

If  /  happens  to  be  nonnegative  on  R,  then,  for  i,  j  =  1 , . . . ,  n,  the  individual 
terms  /(cy )  A  Ay  in  S  may  be  considered  to  be  volumes  of  boxes  having  base  area 


Chapter  5  |  Multiple  Integration 


z 


Figure  5.11  The  volume  under  the  graph  Figure  5.12  The  Riemann  sum  as  a  signed  sum  of 

of  /  is  approximated  by  the  Riemann  sum.  volumes  of  boxes. 


Ax,  Ay j  and  height  /(cy ).  Therefore,  S  can  be  considered  to  be  an  approximation 
to  the  volume  under  the  graph  of  /  over  R,  as  suggested  by  Figure  5.1 1.  If  /  is 
not  necessarily  nonnegative,  then  the  Riemann  sum  S  is  a  signed  sum  of  such 
volumes  (because,  with  /(cy)  <  0,  the  term  /(Cij)AA,j  is  the  negative  of  the 
volume  of  the  appropriate  box — see  Figure  5.12). 


Al\ 

Figure  5.13  HAUA2,A3 
represent  the  values  of  the  shaded 
areas,  then 

f*  f(x)dx  =  Al-A2  +  A3. 


R 

(in  xy-plane) 

Figure  5.14  If  V\,  V2  represent 
the  volumes  of  the  shaded  regions, 
then  ffR  f(x,  y)d A  =  Vl-V2. 


DEFINITION  2.3  The  double  integral  of  /  on  R,  denoted  by  ffRf  dA 
(or  by  f fR  f{x,  y)  dA  or  by  / fR  f(x,  y)  dx  dy),  is  the  limit  of  the  Riemann 
sum  S  as  the  dimensions  Ax;  and  Ayj  of  the  subrectangles  Rjj  all  approach 
zero,  that  is, 

f  f  fdA  =      lim      V  f(Cij)AxiAyj, 

provided  of  course,  that  this  limit  exists.  When  ffRf  dA  exists,  we  say  that 
/  is  integrable  on  R. 


The  crucial  idea  to  remember — indeed,  the  defining  idea — is  that  the  integral 
ffRf  dA  is  a  limit  of  Riemann  sums  S,  for  this  concept  is  what  is  needed  to 
properly  apply  double  integrals  to  physical  situations. 

From  a  geometric  point  of  view,  just  as  the  single-variable  definite  integral 
fa  f(x)  dx  can  be  used  to  compute  the  "net  area"  under  the  graph  of  the  curve 
v  =  f(x)  (as  in  Figure  5. 13),  the  double  integral  ffR  fdA  can  be  used  to  compute 
the  "net  volume"  under  the  graph  of  z  =  fix,  y)  (as  in  Figure  5.14). 

Another  way  to  view  the  double  integral  ffRf  dA  is  somewhat  less  geometric 
but  is  more  in  keeping  with  the  notion  of  the  integral  as  the  limit  of  Riemann 
sums  and  provides  a  perspective  that  generalizes  to  triple  integrals  of  functions  of 
three  variables.  Instead  of  visualizing  the  graph  of  z  =  fix,  y)  as  a  surface  and 
S  =  Yll  j=\  f(cij)AxiAyj  as  a  (signed)  sum  of  volumes  of  boxes  related  to  the 
graph,  consider  S  to  be  a  weighted  sum  of  areas  and  the  integral  ffRf  dA  the 
limiting  value  of  such  weighted  sums  as  the  dimensions  of  all  the  subrectangles 
approach  zero.  With  this  point  of  view,  we  do  not  depict  the  integrand  /  when  we 
try  to  visualize  the  integral.  In  this  way,  the  distinction  between  the  roles  of  the 
integrand  and  the  rectangle  R  over  which  we  integrate  can  be  made  clearer.  (See 
Figure  5.15.) 


5.2  |  Double  Integrals  317 


R 


-  This  subrectangle  of 
area  A^453  contributes 
/(c53)  AA53  to  S. 


Figure  5.15  S  =  /(Cy)AAy. 


EXAMPLE  1  Suppose  that  a  3  cm  square  metal  plate  is  made,  but  some  nonuni- 
formities  exist  due  to  the  manufacturing  process  so  that  the  mass  density  varies 
somewhat  throughout  the  plate.  If  we  knew  the  density  function  8(x,  y)  at  every 
point  in  the  plate,  then  we  could  calculate  the  total  mass  of  the  plate  as 

Total  mass  =  j j  S(x,y)dA, 

where  D  denotes  the  square  region  of  the  plate  placed  in  an  appropriate  coordinate 
system. 

In  the  absence  of  an  analytic  expression  for  S ,  we  nonetheless  can  approximate 
the  double  integral  by  means  of  a  Riemann  sum:  We  partition  the  square  region 
of  the  plate,  take  density  readings  at  a  test  point  in  each  subregion,  and  combine 
to  approximate  the  integral  for  the  total  mass.  (Essentially  what  we  are  doing  is 
assuming  that  the  density  is  nearly  constant  on  each  subregion  so  that  multiplying 
density  and  area  will  give  the  approximate  mass  of  the  subregion;  adding  these 
approximate  masses  then  gives  an  approximation  for  the  total  mass.)  For  example, 
we  might  model  the  problem  as  in  Figure  5.16,  where  the  region  of  the  plate  is 


(0.6) 


(0.3) 


(0.2) 


(0.2) 


(0.1) 


(0.5) 


(0.3) 


(0.3) 


Figure  5.16  The  region  of  Example  1.  The  3x3 
square  is  partitioned  into  nine  subregions.  The  density 
values  at  test  points  in  each  subregion  are  shown. 


Chapter  5  |  Multiple  Integration 


Figure  5.17  The  graph 
of  z  =  x  of  Example  2. 


HI 


Figure  5.1 8  The  two 

subrectangles  R^j  and  Ri2j  are 
symmetrically  placed  with  respect 
to  the  y-axis.  The  corresponding 
test  points  C;,  /  and  c>2 j  are  chosen 
so  that  they  have  the  same 
y-coordinates  and  opposite 
x-coordinates. 


partitioned  into  nine  square  subregions.  Then  we  have 
Total  mass  =  j  j  S(x,  y)dA  «a  ^  5(c,7)AAy 

=  (0.2)1  +  (0.3)1  +  (0.6)1  +  (0.1)1  +  (0.2)1  +  (1)1  +  (0.3)1 

+  (0.3)1 +(0.5)1  =  3.5.  ^ 

EXAMPLE  2  We  determine  the  value  of  ffRxdA,  where  R  =  [-2,2]  x 
[—1,  3].  Here  the  integrand  f(x,  y)  =  x  and,  if  we  graph  z  =  f(x,  y)  over  R, 
we  see  that  we  have  a  portion  of  a  plane,  as  shown  in  Figure  5.17.  Note  that  the 
portion  of  the  plane  is  positioned  so  that  exactly  half  of  it  lies  above  the  xy-plane 
and  half  below.  Thus,  if  we  regard  ffRx  dA  as  the  net  volume  under  the  graph  of 
z  =  x,  then  we  conclude  that  ffRx  dA  (if  it  exists)  must  be  zero. 

On  the  other  hand,  we  need  not  resort  to  visualization  in  three  dimensions. 
Consider  a  Riemann  sum  corresponding  to  ffRx  dA  obtained  by  partitioning 
R  =  [—2,  2]  x  [—1,3]  symmetrically  with  respect  to  the  y-axis  and  by  choosing 
the  "test  points"  c,;  symmetrically  also.  (See  Figure  5.18.)  It  follows  that  the  value 
of 

S  =      f(cu)AAij  =  J2XiiAAiJ 

(where  x,7  denotes  the  x-coordinate  of  cy)  must  be  zero  since  the  terms  of  the 
sum  cancel  in  pairs.  Furthermore,  we  can  arrange  things  so  that,  as  we  shrink 
the  dimensions  of  the  subrectangles  to  zero  (as  we  must  do  to  get  at  the  integral 
itself),  we  preserve  all  the  symmetry  just  described.  Hence,  the  limit  under  these 
restrictions  will  be  zero,  and  thus,  the  overall  limit  (where  we  do  not  impose  such 
symmetry  restrictions  on  the  Riemann  sum),  if  it  exists  at  all,  must  be  zero  as 
well.  ♦ 

Example  2  points  out  fundamental  difficulties  with  Definition  2.3,  namely, 
that  we  never  did  determine  whether  ffRf  dA  really  exists.  To  do  this,  we  would 
have  to  be  able  to  calculate  the  limit  of  Riemann  sums  of  /  over  all  possible 
partitions  of  R  by  using  all  possible  choices  for  the  test  points  c,y,  a  practically 
impossible  task.  Fortunately,  the  following  result  (which  we  will  not  prove)  pro- 
vides an  easy  criterion  for  integrability: 


Figure  5.1 9  The  graph  of  a 
piecewise  continuous  function. 


THEOREM  2.4  If  /  is  continuous  on  the  closed  rectangle  R,  then  ffR  f  dA 
exists. 

In  Example  2,  f(x,  y)  =  x  is  a  continuous  function  and  hence  integrable 
by  Theorem  2.4.  The  symmetry  arguments  used  in  the  example  then  show  that 
ffRxdA  =  0, 

Continuous  functions  are  not  the  only  examples  of  integrable  functions.  In 
the  case  of  a  function  of  a  single  variable,  piecewise  continuous  functions  are  also 
integrable.  (Recall  that  a  function  f(x)  is  piecewise  continuous  on  the  closed 
interval  [a,  b]  if  /  is  bounded  on  [a,  b]  and  has  at  most  finitely  many  points  of 
discontinuity  on  the  interior  of  [a ,  b] .  Its  graph,  therefore,  consists  of  finitely  many 
continuous  "chunks"  as  shown  in  Figure  5.19.)  For  a  function  of  two  variables, 
there  is  the  following  result,  which  generalizes  Theorem  2.4. 


5.2  |  Double  Integrals  319 


THEOREM  2.5  If  /  is  bounded  on  R  and  if  the  set  of  discontinuities  of  /  on  R 
has  zero  area,  then  ffR  f  dA  exists. 


To  say  that  a  set  X  has  zero  area  as  we  do  in  Theorem  2.5,  we  mean  that  we 
can  cover  X  with  rectangles  R\,  R2,  . . . ,  Rni . . .  (i.e.,  so  that  X  c  [J^Lj  ^«)>  tne 
sum  of  whose  areas  can  be  made  arbitrarily  small. 

A  function  /  satisfying  the  hypotheses  of  Theorem  2.5  has  a  graph  that 
looks  roughly  like  the  one  in  Figure  5.20.  Theorem  2.5  is  the  most  general  suffi- 
cient condition  for  integrability  that  we  will  consider.  It  is  of  particular  use  to  us 
when  we  define  the  double  integral  of  a  function  over  an  arbitrary  region  in  the 
plane. 


a  a 


 X 

Discontinuities  of/ 


Figure  5.20  The  graph  of  an  integrable 
function. 


Although  Theorems  2.4  and  2.5  make  it  relatively  straightforward  to  check 
that  a  given  integral  exists,  they  do  little  to  help  provide  the  numerical  value  of 
the  integral.  To  mechanize  the  evaluation  of  double  integrals,  we  will  use  the 
following  result: 


THEOREM  2.6  (Fubini's  theorem)  Let  /  be  bounded  on  R  =  [a,  b]  x  [c,  d] 
and  assume  that  the  set  S  of  discontinuities  of  /  on  R  has  zero  area.  If  every  line 
parallel  to  the  coordinate  axes  meets  S  in  at  most  finitely  many  points,  then 

p  p  pb    pd  pd  pb 

//  fdA=  /    /    f(x,y)dydx=  /     /  f(x,y)dxdy. 

J  J  R  J a    J  c  J  c    J  a 


Fubini's  theorem  demonstrates  that  under  certain  assumptions  the  double  in- 
tegral over  a  rectangle  (i.e.,  the  limit  of  Riemann  sums)  can  be  calculated  by  using 
iterated  integrals  and,  moreover,  that  the  order  of  integration  for  the  iterated  inte- 
gral does  not  matter.  We  remark  that  the  independence  of  the  order  of  integration 
depends  strongly  on  the  fact  that  the  region  of  integration  is  rectangular;  it  will  not 


320       Chapter  5  |  Multiple  Integration 


generalize  to  more  arbitrary  regions  in  such  a  simple  way.  (A  proof  of  Theorem 
2.6  is  given  in  the  addendum  to  this  section.) 

EXAMPLE  3  Werevisit  ffRx  dA  in  Example  2,  where  R  =  [-2,  2]  x  [-1,  3]. 
By  Theorem  2.6,  we  know  that  JfRxdA  exists  and  by  Fubini's  theorem,  we 
calculate 


j  j  xdA  =  j  j  x  dy  dx  =  j 


xy 


=-i 


dx 


-L 


x(3  -  {-\))dx 


/: 


Ax  dx  =  2x 


2|2 

1-2 


8  =  0, 


which  checks.  Furthermore,  we  also  have 


j  j  x  dA  =  j     j    xdxdy  =  j 

J  JR  J-l  J-2  J-] 


dy  =  f_l(2-2)dy  =  0. 

x=—2 


PROPOSITION  2.7  (Properties  of  the  integral)  Suppose  that  /  and  g  are 
both  integrable  on  the  closed  rectangle  R.  Then  the  following  properties  hold: 


1.  /  +  g  is  also  integrable  on  R  and 


/  /  (f  +  g)dA=  f  f  fdA+  f  f  gdA. 
J  Jr  J  J r  J  J r 


2.  cf  is  also  integrable  on  R,  where  c  e  R  is  any  constant,  and 


\LcfiA  =  c\lSdA' 


3.  If  f(x,  y)  <  g(x,  y)  for  all  (x,y)e  R,  then 


If  f{x,y)dA<  I j  g(x,y)dA. 


4.  |  f\  is  also  integrable  on  R  and 


f Ir  f Jr 


<  /  /  \f\dA. 


Properties  1  and  2  are  called  the  linearity  properties  of  the  double  integral. 
They  can  be  proved  by  considering  the  appropriate  Riemann  sums  and  taking 
limits.  For  example,  to  prove  property  1,  note  that  the  Riemann  sum  whose  limit 


5.2  |  Double  Integrals  321 


is 


ffR(f  +  g)dA 


is 


Figure  5.21 

in  the  plane. 


A  bounded  region  D 


Figure  5.22  A  type  1  elementary 


region. 

) 

> 

y  =  d 

POO 

y  =  c 

Figure  5.23  A  type  2  elementary 
region. 


t'J=i 


=  E  /(cv)AA!7  +  E  *(c/y)AAy 

U    I  1,7  =  1 


Property  3  (known  as  monotonicity)  and  property  4  can  also  be  proved  using 
Riemann  sums.  For  property  4,  one  needs  to  use  the  fact  that 


E«* 

k=l 


<E|a*!- 

fc=l 


Double  Integrals  over  General  Regions  in  the  Plane  

Our  next  step  is  to  understand  how  to  define  the  integral  of  a  function  over  an 
arbitrary  bounded  region  D  in  the  plane.  Ideally,  we  would  like  to  give  a  precise 
definition  of  ffDfdA,  where  D  is  the  amoeba-shaped  blob  shown  in  Figure  5.21 
and  where  /  is  bounded  on  D.  In  keeping  with  the  definition  of  the  integral  over 
a  rectangle,  ffD  f  dA  should  be  a  limit  of  some  type  of  Riemann  sum  and  should 
represent  the  net  volume  under  the  graph  of  /  over  D.  Unfortunately,  the  techni- 
calities involved  in  making  such  a  direct  approach  work  are  prohibitive.  Instead,  we 
shall  consider  only  certain  special  regions  (rather  than  entirely  arbitrary  ones),  and 
we  shall  assume  that  the  integrand  /  is  continuous  over  the  region  of  integration 
(which  will  allow  us  to  use  what  we  already  know  about  integrals  over  rectangles). 
Although  this  approach  will  not  provide  us  with  a  completely  general  definition, 
it  is  sufficient  for  essentially  all  the  practical  situations  we  will  encounter. 
To  begin,  we  define  the  types  of  elementary  regions  we  wish  to  consider. 


DEFINITION  2.8  We  say  that  D  is  an  elementary  region  in  the  plane  if  it 
can  be  described  as  a  subset  of  R2  of  one  of  the  following  three  types: 

Type  1  (see  Figure  5.22): 

D  =  {(*,  y)  |  y(x)  <y<  S(x),  a<x  <b}, 
where  y  and  <5  are  continuous  on  [a,  b]. 
Type  2  (see  Figure  5.23): 

D  =  {(x,  y)  |  a(y)  <  x  <  0(y),  c<y<d}, 
where  a  and  /3  are  continuous  on  [c,  d]. 
Type  3  D  is  of  both  type  1  and  type  2. 


Thus,  a  type  1  elementary  region  D  has  a  boundary  (denoted  3D)  consisting 
of  straight  segments  (possibly  single  points)  on  the  left  and  on  the  right  and  graphs 
of  continuous  functions  of  x  on  the  top  and  on  the  bottom.  A  type  2  elementary 


322       Chapter  5  |  Multiple  Integration 


region  has  a  boundary  that  is  straight  on  the  top  and  bottom  and  consists  of  graphs 
of  continuous  functions  of  y  on  the  left  and  right. 

EXAMPLE  4  The  unit  disk,  shown  in  Figure  5.24,  is  an  example  of  a  type  3 
elementary  region.  It  is  a  type  1  region  since 


D  =  Ux,y)  |  — v/l  -  x2 


y  <  yi  — . 

(See  Figure  5.25.)  It  is  also  a  type  2  region  since 


D  =  Ux,  y)  |  —/l  -  y2  <x< 
(See  Figure  5.26.) 


y 


-1  <x  <  lj  . 

i<y<i}- 


Figure  5.24  The  unit  disk 
D  =  {(x,  y)  |  a2  +  v2  <  1}  is  a 
type  3  region. 


y  =  l 

/ 

y  =  -l 

Figure  5.25  The  unit  disk  D  as  a 
type  1  region. 


Figure  5.26  The  unit  disk  D  as  a  type  2 
region. 


Now  we  are  ready  to  define  ffDf  dA,  where  D  is  an  elementary  region  and 
/  is  continuous  on  D.  We  construct  a  new  function  /ext,  the  extension  of  /,  by 


fex\x,y)  = 


f(x,y)    if  (x,y)eD 
0  if(x,y)£D 


Note  that,  in  general,  /ext  will  not  be  continuous,  but  the  discontinuities  of  /ext 
will  all  be  contained  in  3D,  which  has  no  area.  Hence,  by  Theorem  2.5,  /ext  is 
integrable  on  any  closed  rectangle  R  that  contains  D.  (See  Figure  5.27.) 


Figure  5.27  The  graph  of  z  =  /ext(jc,  y). 


5.2  |  Double  Integrals 


jy  =  3x2 

Djl(l,3) 

/       \  y  =  4  -  X2 

(2,0) 

Figure  5.28  The  domain  of  /  of 
Example  5. 


DEFINITION  2.9  Under  the  previous  assumptions  and  notation,  if  R  is  any 
rectangle  that  contains  D,  we  define 


fLfdA  ,obe  //« 


/extJA. 


Note  that  Definition  2.9  implicitly  assumes  that  the  choice  of  the  rectangle  R 
that  contains  D  does  not  affect  the  value  of  f fR  fext  dA.  This  is  almost  obvious 
but  still  should  be  proved.  We  shall  not  do  so  directly  but  instead  establish  the 
following  key  result: 

THEOREM  2.10  Let  D  be  an  elementary  region  in  R2  and  /  a  continuous 
function  on  D. 

1.  If  D  is  of  type  1  (as  described  in  Definition  2.8),  then 

/  /  fdA  =  /     /  f(x,y)dydx. 

J  J D  Ja  Jy(x) 

2.  If  D  is  of  type  2,  then 

r  r  rd  pP(y) 

fdA=        /  f(x,y)dxdy. 

J  J  D  Jc  Ja(y) 


Theorem  2.10  provides  an  explicit  and  straightforward  way  to  evaluate  double 
integrals  over  elementary  regions  using  iterated  integrals.  Before  we  prove  the 
theorem,  let  us  illustrate  its  use. 

EXAMPLE  5  Let  D  be  the  region  bounded  by  the  parabolas  y  =  3x2,  y  = 
4  —  x2  and  the  y-axis  as  shown  in  Figure  5.28.  (Note  that  the  parabolas  intersect  at 
thepoint(l,  3).)  Since  D  isatype  1  elementary  region,  we  may  use  Theorem  2. 10 
with  f(x,  y)  —  x2y  to  find  that 


2       fl  r4^2  2 

x  ydA=  II       x  ydydx. 
Jo  J\k2 


The  limits  for  the  first  (inside)  integration  come  from  the  y-values  of  the  top  and 
bottom  boundary  curves  of  D.  The  limits  for  second  (outside)  integration  are  the 
constant  x -values  that  correspond  to  the  straight  left  and  right  sides  of  D.  The 
evaluation  itself  is  fairly  mechanical: 

=4-x2 


I     I       x2  ydydx  =  I  I 

J0    Jxx2  Jo  V 


1/22 

'  x  y 


dx 


=3x2 


=  /  y((4-*2)2-(3*2)>* 
l  fl 

=  -J   x2  (16  -  8;t2  +  x4  -  9x4)  dx 

=  f  (8.t2  -  4x4  -  4x6)  dx  = 
Jo 


8  4  4  _  136 
3       5       7  105- 


324       Chapter  5  |  Multiple  Integration 


Note  that  after  the  y -integration  and  evaluation,  what  remains  is  a  single  definite 
integral  in  x.  The  result  of  calculating  this  x -integral  is,  of  course,  a  number.  Such 
a  situation  where  the  number  of  variables  appearing  in  the  integral  decreases  with 
each  integration  should  always  be  the  case.  ♦ 


Proof  of  Theorem  2.10  For  part  1,  we  may  take  D  to  be  described  as 

D  =  {(x,  y)  |  y(x)  <  y  <  S(x),  a  <  x  <b}. 
We  have,  by  Definition  2.9,  that 


fLfdA=fLf~dA- 


where  R  is  any  rectangle  containing  D.  Let  R  =  [a',  b']  x  [c',  d'],  where  a'  <  a, 
b'  >  b,  and  c'  <  y(x),  d'  >  S(x)  for  all  x  in  [a,  b].  That  is,  we  have  the  situation 
depicted  in  Figure  5.29.  Since  /ext  is  zero  outside  of  the  subrectangle  R2  = 
[a,b]  x  [c',d'], 

f  f  fextdA=  f  f  fextdA=  f    f  fex\x,y)dydx 

by  Fubini's  theorem.  For  a  fixed  value  of  x  between  a  and  b,  consider  the  y- 
integral  fext(x,  y)dy.  Since  fext(x,  y)  =  0 unless  y(x)  <  y  <  S(x)  (in  which 
case  fext(x,  y)  =  f(x,  y)), 


pel'  pS(x) 

/    rx\x,y)dy=  /  f(x,y)dy, 

Jd  Jy(x) 


and  so 


f  f  f(x,y)dA=  f  f  fxtdA=  f    f  fex\x,y)dydx 

J  JD  J  J R  Ja  Jc' 


/    /  f(x,y)dydx, 


as  desired. 

The  proof  of  part  2  is  very  similar. 


Figure  5.29  The  region  R  is  the  union  of  Ri,  R2, 
and^?3 . 


(0,1)' 

>v x  +  y  =  1 

D 

(0,0) 

\(1.0) 

Figure 

5.30  The  region  D  of 

Example  6. 

; 

> 

\  J  =  1  -  X  | 

x  =  0 

D 

y  =  0 

Figure 

5.31  The  region  D  of 

Example  6  as  a  type  1  region. 

; 

> 

y  =  l 

x  =  0 

\x  =  1  —  y 

D  \. 

y  =  0 

Figure 

5.32  The  region  D  of 

Example  6  as  a  type  2  region. 


5.2  |  Double  Integrals  325 

We  continue  analyzing  examples  of  double  integral  calculations. 

EXAMPLE  6  Let  D  be  the  region  shown  in  Figure  5.30  having  a  triangular 
border.  Consider  f fD(\  —  x  —  y)dA.  Note  that  D  is  a  type  3  elementary  region, 
so  there  should  be  two  ways  to  evaluate  the  double  integral. 

Considering  D  as  a  type  1  elementary  region  (see  Figure  5.3 1),  we  may  apply 
part  1  of  Theorem  2.10  so  that 

/  /  (1  —  x  —  y)dA=  I    I      (1  —  x  —  y)dydx 
J  3d  Jo  Jo 


-L 
-L 
-I 


y-xy 


y=\-x 


dx 


y=0 


'  7                              (1  -x)2\ 
(1  —  x)  —  x(l  —  x)  I  dx 


(1  -  xf 


dx=  -\{\-x)\  =  \. 


We  can  also  consider  D  as  a  type  2  elementary  region,  as  shown  in  Figure  5.32. 
Then,  using  part  2  of  Theorem  2.10,  we  obtain 

/  /  (1  —  x  —  y)dA  =  I     I  (l-x-y)dxdy. 

J  J D  Jo  Jo 

We  leave  it  to  you  to  check  explicitly  that  this  iterated  integral  also  has  a  value  of 
t  .  Instead  we  note  that 


Jo  Jo 


y  )  dy  dx 


can  be  transformed  into 


/     /      (1  —  x  —  y)  dx  dy 
Jo  Jo 

by  exchanging  the  roles  of  x  and  y.  Hence,  the  two  integrals  must  have  the  same 
value.  In  any  case,  the  double  integral 


(1  —  x  —  y)dA 


represents  the  volume  under  the  graph  of  z  =  1  —  x  —  y  over  the  triangular  region 
D .  If  we  picture  the  situation  in  R3 ,  as  in  Figure  5 . 3  3 ,  we  see  that  the  double  integral 
represents  the  volume  of  a  tetrahedron.  ♦ 


Of  course,  not  all  regions  in  the  plane  are  elementary,  including  even  some 
relatively  simple  ones.  To  integrate  continuous  functions  over  such  regions,  the 
best  advice  is  to  attempt  to  subdivide  the  region  into  finitely  many  of  elementary 
type. 


326       Chapter  5  |  Multiple  Integration 


EXAMPLE  7  Let  D  be  the  annular  region  between  the  two  concentric  circles 
of  radii  1  and  2  shown  in  Figure  5.34.  Then  D  is  not  an  elementary  region,  but  we 
can  break  D  up  into  four  subregions  that  are  of  elementary  type.  (See  Figure  5.35.) 
If  f(x,  y)  is  any  function  of  two  variables  that  is  continuous  (hence  integrable) 
on  D,  then  we  may  compute  the  double  integral  as  the  sum  of  the  integrals  over 
the  subregions.  That  is, 

[[  fdA=[[  fdA+ff  fdA+ff  fdA+[[  fdA. 

J  J D  J  J Di  J  JD2  J  J D]  J  J D4 

For  the  type  1  subregions,  we  have  the  set-up  shown  in  Figure  5.36: 

f(x,  y)dy  dx 


and 


\LSiA 


V3  ,-1 


f(x,  y)dy  dx. 


-VI  J—jA^? 

For  the  type  2  subregions,  we  use  the  set-up  shown  in  Figure  5.37: 
f  fo2  f-i  f*/i 


y  =  — V4  -  x- 


Figure  5.36  The  subregions  D\  and 
Z?3  of  Example  7  are  of  type  1 . 


fix,  y)dx  dy 

J 

1 

.2  _ 

\  ✓ 

\ 

\ 

s 

X 

✓ 

x=  VT^y2 


x  =  V4-y2 


Figure  5.37  The  subregions  D2  and  D4  of 
Example  7  are  of  type  2. 


5.2  |  Double  Integrals 


and 


JL<«-U. 


<l-y2 


-1  J-J4-V2 


f(x,  y)dx  dy. 


The  difficulty  of  evaluating  each  of  the  preceding  four  iterated  integrals  then 
depends  on  the  complexity  of  the  integrand.  ♦ 


2 

/  D 

/x-y=0 

~K  ~l  / 

1  2 

-l 

Figure  5.38  The  region  D 
of  Example  8. 


EXAMPLE  8  We  calculate  ffD  ydA,  where  D  is  the  region  bounded  by  the 
line  x  —  y  =  0  and  the  parabola  x  =  y2  —  2.  (See  Figure  5.38.) 

In  this  case  D  is  a  type  2  elementary  region,  where  the  left  and  right  boundary 
curves  may  be  expressed  as  x  =  y2  —  2  and  x  =  y,  respectively.  These  curves 
intersect  where 


y  =  -l,2. 


y*-2  =  y  y^-y-2  =  Q  < 

Therefore,  part  2  of  Theorem  2.10  applies  to  give 

-2  py 


1L 


ydA 


/I  y  dx  dy 
■1  Jv2-2 


f  i  f2  2  xy\^_2  dy  =  j  y-  (y2  -  2)y)  dy 
j  y-y'+2y)  dy- 


3  4  \  ^ 

y--y-  +  y2 
3      4  y 


=  (f-4  +  4)-(- 


3  4 


+  !)  =  !■ 


Now,  although  D  is  not  a  type  1  elementary  region,  it  may  be  divided  into 
two  type  1  subregions  along  the  vertical  line  x  =  —  1.  (See  Figure  5.39.)  The 
subregion  D\  lying  left  of  the  line  x  =  —  1  is  bounded  on  both  top  and  bottom  by 
the  parabola  x  =  y2  —  2;  by  solving  for  y  we  may  express  the  bottom  boundary 
of  D\  as  y  =  —*Jx  +  2  and  the  top  boundary  as  y  =  \fx  +  2.  The  subregion 
D2  lying  right  of  x  =  —  1  is  bounded  on  the  bottom  by  y  =  x  and  on  the  top  by 


Figure  5.39  The  region  D  of  Example  8  is  divided 
into  subregions  D\  and  D2  by  the  line  x  =  —  1. 


328       Chapter  5  |  Multiple  Integration 


y  =  -J  x  +  2.  Putting  all  this  information  together,  we  have 
/  /  ydA  =  I  I    ydA+  f  f  ydA 


/ —  1    p*/x+2  r2  r 

I  _ydydx+  /  / 
-2    J-Jx+2  J -I  Jx 


y  dy  dx 


-1  y2 

<-2  ~2 

-1 


y=Vx+2 


dx  + 


£ 


2    2  y=vx+2 


y=-Vx+2 

/■2/x  +  2  *2\ 

-  y  ( — 


,  x3 
1  4+*"T 


=  (l+2-l)-(i-l  +  i) 


4' 


Addendum:  Proof  of  Theorem  2.6   

Step  1.  First  we  establish  Theorem  2.6  in  the  case  where  /  is  continuous  on 
R  =  [a,  b]  x  [c,  d].  By  Theorem  2.4,  we  know  that  ffRf  dA  exists.  Let  F  be 
the  single-variable  function  denned  by 


F(x) 


f(x,y)dy. 


(Note:  Since  /  is  continuous  on  R,  the  partial  function  in  y  obtained  by  holding 
x  constant  is  continuous  on  [c,  d].  Hence,  fd  fix,  y)dy  exists  for  every  x  in 
[a,  b].)  We  show  that 

nb  nb  r    nd  n  n 

/   F(x)dx=  /      /    f(x,y)dy   dx  =        f  dA. 

J  a  J  a     \_Jc  J  J  JR 

Let  a  =  xq  <  x\  <  •  •  •  <  x„  =  b  be  any  partition  of  [a,  b].  Then  a  general 
Riemann  sum  that  approximates     F(x)  dx  is 

// 

j>(jtf)A*,,  (1) 


where  Ax;  =  x;-  —  andx*  e  [x,-_i,  x,].Nowletc  =  y0  <  <  •  •  •  <  =  d 
be  a  partition  of  [c,  d].  (The  partitions  of  [a,  b]  and  [c,  J]  together  give  a  partition 
of  R  =  [a,  b]  x  [c,  rf].)  Therefore,  we  may  write 


x)  =  Jdf(x, 


y)dy 


fix,  y)dy  + 


y)dy  + 


t  f 

7=1  Jy>-1 


f(x,y)dy. 


+  f  fix,y)dy 

Jy„-i 


5.2  |  Double  Integrals  329 

By  the  mean  value  theorem  for  integrals,1  on  each  subinterval  [yj-i,  yj]  there 
exists  a  number  y*  such  that 


f 

Jy, 


f(x,  y)dy  =  (yj  -  y}-i)f(x,  y*)  =  f(x,  y*)Ayj. 


The  choice  of  y*  in  general  depends  on  x,  so  henceforth,  we  will  write  y*(x)  for 
y*.  Consequently, 

n 

/•(.v)- ^./(.v.y;(.v)).\y,. 


and  the  Riemann  sum  (1)  may  be  written  as 


i=l  I  7=1 


Axi  =  /(C;,).\.V,.\.V(. 


where  cy  =  (xf ,  y*(xf)).  Note  that  Cy  e  x;]  x  [y;_i,  y,-].  (See  Fig- 
ure 5.40.) 


y;(*,*) 


c  • 

H  1 — h 


H  h 


 1    «  I 


Figure  5.40  The  point  Cy  =  (.«*,  y  *(*,*))  used  in 
the  proof  of  Theorem  2.6. 

We  have  thus  shown  that  given  any  partition  of  [a,b],  we  can  associate  a 
suitable  partition  of  R  =  [a,  b]  x  [c,  d]  such  that  the  Riemann  sum  (1)  that  ap- 
proximates fa  F(x)dx  is  equal  to  a  Riemann  sum  (namely,  J2i  j  f(cij)Axj  Ayj) 
that  approximates  ffRf  dA.  Since  /  is  continuous,  we  know  that 

/(cy)Ajc,- Ay7-    approaches     /  /  f  dA 
U  J  Js 

as  Ax,  and  Ay,-  tend  to  zero.  Hence, 

I  F(x)dx  =  11  fdA. 

Ja  J  JR 

By  exchanging  the  roles  of  x  and  y  in  the  foregoing  argument,  we  can  show 

that 

II fdA =  f  I  f(x,y)dxdy. 

J  J  R  Jc  Ja 


1  The  mean  value  theorem  for  integrals  says  that,  if  g  is  continuous  on  [a,  b],  then  there  is  some  number 
c  with  a  <  c  <  b  such  that  f  g(x)dx  =  (b  —  a)g(c). 


Chapter  5  |  Multiple  Integration 


Step  2.  Now  we  prove  the  general  case  of  Theorem  2.6  (i.e.,  the  case  that 
/  has  discontinuities  in  R  =  [a,  b]  x  [c,  d]).  By  hypothesis,  the  set  S  of  discon- 
tinuities of  /  in  R  are  such  that  every  vertical  line  meets  S  in  at  most  finitely 
many  points.  Thus,  for  each  x  in  [a,  b],  the  partial  function  in  y  of  f(x,  y)  is 
continuous  throughout  [c,  d],  except  possibly  at  finitely  many  points.  (In  other 
words,  the  partial  function  is  piecewise  continuous.)  Then,  because  /  is  bounded, 


(x)  =  ^  f(x 


F(x)=  I  f(x,y)dy 
exists. 

Now  we  proceed  as  in  Step  1 .  That  is,  we  begin  with  a  partition  of  [a,  b]  into 
n  subintervals  and  a  corresponding  Riemann  sum 

i=i 

Next,  we  partition  [c,  d]  into  n  subintervals.  Hence, 

Fix*)  =  f  f(x*,  y)dy  =  J2  f'  fix*,  y)dy.  (2) 

As  in  Step  1,  the  partitions  of  [a,  b]  and  [c,  d]  combine  to  give  a  partition  of  R. 
Write  R  as     U  R2,  where  R\  is  the  union  of  all  subrectangles 

Rij  =  [Xi-uXf]  x  [y;_!,  yj] 

that  intersect  S  and  R2  is  the  union  of  the  remaining  subrectangles.  Then  we  may 
apply  the  mean  value  theorem  for  integrals  to  those  intervals  [yj-i,  yj]  on  which 
fix*,  y)  is  continuous  in  y,  thus  replacing  the  integral 

/  f(xf,y)dy 

by 

f(x*,y*(x*))Ayj  =  f(cij)Ayj. 

Since  /  is  bounded,  we  know  that 

\f(x,y)\<M 

for  some  M  and  all  (x,  y)  e  R.  Therefore,  on  the  intervals  [y7_i,  yj]  where 
fix* ,  y)  fails  to  be  continuous,  we  have 

f1  fix*,y)dy    <  r  \  f(x*,y)\dy 

<  r  Mdy  =  M(yj  -  y;_j)  =  MAyr  (3) 

Jyj-i 

From  equation  (2),  we  know  that 

f2Hx*)Axi  =  it,\f}  fix*,y)dy}AXi 

=    E    f  fix*,y)dy\AXi 


5.2  |  Double  Integrals  331 


f(x*,y)dy\Axi 


f(x*,y)dy\Ax 


Therefore, 


= s  \i: 

A',,t  K-  1  *'  '•  • 

J>(x*)Ax,-  J]  {  r  f(xf,y)dy\Axi 

-  (f 


f(x*,y)dy^Axi 


(4) 


Applying  the  mean  value  theorem  for  integrals  to  the  left  side  of  equation  (4)  and 
inequality  (3)  to  the  right  side,  we  obtain 


J2nx*)Axi-  f(Cij)AxtAyj 

i  =  l  RijCRi 


MAxjAyj 

RijCRi 

M  ■  area  of  R\ . 


Now  S  has  zero  area  (by  hypothesis)  and  is  contained  in  R\.  By  letting  the 
partition  of  R  become  sufficiently  fine  (i.e.,  by  making  Ax, ,  Ay,  small),  the  term 
M  ■  area  of  R\  can  be  made  arbitrarily  small.  (See  Figure  5.41.) 


d  - 

c  - 

- — 

N 

< 

f 

i 

— - 

1 

a 

b 

Figure  5.41  The  set  R\  (shaded  area)  consists  of 
the  subrectangles  of  the  partition  of  R  that  meet  S, 
the  set  of  discontinuities  of  /  on  R.  As  the  partition 
becomes  finer,  the  area  of  Ri  tends  toward  zero. 


Therefore,  as  all  Ax,  and  Ay,-  tend  to  zero,  we  have  that  the  sums 
J2F(xf)Axi    and  /<e»/).Yv,.\.\> 

i  RijCRi 

and  the  term  M  ■  area  of  R\  converge  (respectively)  to 

f  F(x)dx,     f  f  fdA,    and  0. 

J  a  J  JR 


332       Chapter  5  |  Multiple  Integration 

We  conclude  that 

that  is, 


5.2  Exercises 


f  F(x)dx  -  f  f  fdA  =  0, 

J  a  J  J  R 

(  (  fdA  =  f    f  f(x,y)dydx. 

J  J  R  J  a  Jc 

Again,  by  exchanging  the  roles  of  x  and  y,  we  can  show  that 

"d  pb 


as  well. 


[  (  fdA  =  f    f  f(x,y)dxdy 

J  JR  Jc    J  a 


1.  Use  Definition  2.3  and  Theorem  2.4  to  determine 
the  value  of  f fR(y3  +  s'm2y)dA,  where  R  =  [0,  3]  x 
[-1,1]- 

2.  Let  R  =  [-3,  3]  x  [-2,  2].  Without  explicitly  eval- 
uating any  iterated  integrals,  determine  the  value  of 
JfR(x5+2y)dA. 

3.  This  problem  concerns  the  double  integral  ffpX^dA, 
where  D  is  the  region  pictured  in  Figure  5.42. 

(a)  Determine  ff£)xidA  directly  by  setting  up  and  ex- 
plicitly evaluating  an  appropriate  iterated  integral. 

(b)  Now  argue  what  the  value  of  ffD  x3  dA  must  be 
by  inspection,  that  is,  without  resorting  to  explicit 
calculation. 


Figure  5.42  The  region  D  of 
Exercise  3 . 

In  Exercises  4-13,  evaluate  the  given  iterated  integrals.  In  ad- 
dition, sketch  the  regions  D  that  are  determined  by  the  limits 
of  integration. 


,  f    f  3d 

Jo  Jo 


r2  r1 

i.   I     I     y  dx  dy 

Jo  Jo 


Jo  Jo 


lo  Jo 

^3  ?2x+\ 


y  dy  dx 


-1  xydy 


dx 


2  +  y2)dy  dx 


10. 


11. 


12. 


13. 


/  /  {x 

Jo  Jx2/4 
,4  ,2,/> 

/  /  x  sin  iy  )  dx  dy 
Jo  Jo 

pit  psinx 

11       y  cos  x  dy  dx 
Jo  Jo 

f    f       '  3dydx 
Jo  J-JT^x1 

Li  zdxd> 

lo  J-t 


y  dy  dx 


14.  Figure  5.43  shows  the  level  curves  indicating  the  vary- 
ing depth  (in  feet)  of  a  25  ft  by  50  ft  swimming  pool. 
Use  a  Riemann  sum  to  estimate,  to  the  nearest  100  ft3, 
the  volume  of  water  that  the  pool  contains. 


Figure  5.43 

1 5.  Integrate  the  function  f(x,  y)  =  1  —  xy  over  the  trian- 
gular region  whose  vertices  are  (0,  0),  (2,  0),  (0,  2). 


5.2  I  Exercises  333 


16.  Integrate  the  function  f(x,  y)  =  3xy  over  the  region 
bounded  by  y  =  32x3  and  y  =  ~Jx. 

17.  Integrate  the  function  f(x,  y)  =  x  +  y  over  the  region 
bounded  by  x  +  y  =  2  and  y2  —  2y  —  x  =  0. 

18.  Evaluate  ffDxydA,  where  D  is  the  region  bounded 
by  x  =  y3  and  y  =  x2. 

19.  Evaluate  ffD  e*2  d A,  where  D  is  the  triangular  region 
with  vertices  (0,  0),  (1,  0),  and  (1,  1). 

20.  Evaluate  ffD  3yd A,  where  D  is  the  region  bounded 
by  xy2  =  1,  y  =  x,  x  =  0,  and  y  =  3. 

21.  Evaluate  ffD(x  —  2y)dA,  where  D  is  the  region 
bounded  by  y  =  x2  +  2  and  y  =  2x2  —  2. 

22.  Evaluate  JfD(x2  +  y2)dA,  where  D  is  the  region  in  the 
first  quadrant  bounded  by  y  =  x,  y  =  3x,  and  xy  =  3. 

23.  Prove  property  2  of  Proposition  2.7. 

24.  Prove  property  3  of  Proposition  2.7. 

25.  Prove  property  4  of  Proposition  2.7. 

26.  (a)  Let  D  be  an  elementary  region  in  R2.  Use  the 

definition  of  the  double  integral  to  explain  why 
ffD  IdA  gives  the  area  of  D. 

(b)  Use  part  (a)  to  show  that  the  area  inside  a  circle  of 
radius  a  isjia2. 

27.  Use  double  integrals  to  find  the  area  of  the  region 
bounded  by  y  =  x2  and  y  =  x3. 

28.  Use  double  integrals  to  calculate  the  area  of  the  region 
bounded  by  y  =  2x,  x  =  0,  and  y  =  1  —  2x  —  x2. 

29.  Use  double  integrals  to  calculate  the  area  inside 
the  ellipse  whose  semiaxes  have  lengths  a  and  b. 
(See  Figure  5.44.) 


Figure  5.44  The  ellipse  of 
Exercise  29. 


30.  (a)  Set  up  an  appropriate  iterated  integral  to  find 
the  area  of  the  region  bounded  by  the  graphs  of 
y  =  —  x  and  y  =  ax2  for  x  >  0.  (Take  a  to  be 
a  constant.) 

(b)  Use  a  computer  algebra  system  to  estimate  for  what 
value  of  a  this  area  equals  1 . 


31 .  Use  double  integrals  to  find  the  total  area  of  the  region 
bounded  by  y  =  x3  and  x 


y5- 


32.  Use  double  integrals  to  find  the  area  of  the  region 
bounded  by  the  parabola  y  =  2  —  x2,  and  the  lines 
X  -  y  =  0,  2x  +  y  =  0. 

33.  Let  D  be  the  region  in  R2  bounded  by  the  lines  x  =  0, 
x  +  y  =  3,  and  x  —  y  =  3.  Without  resorting  to  any 
explicit  calculation  of  an  iterated  integral,  determine, 
with  explanation,  the  value  of  ffD(y3  —  ex  siny  + 
2)dA.  (Hint:  Use  symmetry  and  geometry.) 

34.  Let  D  be  the  region  in  R2  with  y  >  0  that  is 
bounded  by  x2  +  y2  =  9  and  the  line  y  =  0.  Without 
resorting  to  any  explicit  calculation  of  an  iterated 
integral,  determine,  with  explanation,  the  value  of 


ffDV* 


y  sin*  —  2)  dA. 


35.  Determine  the  volume  of  the  solid  lying  under  the  plane 
z  =  24  —  2x  —  6y  and  over  the  region  in  the  jcy-plane 
bounded  by  y  =  4  —  x2,  y  =  4x  —  x2,  and  the  y-axis. 


36.  Find  the  volume  under  the  portion  of  the  paraboloid 
z  =  x2  +  6y2  lying  over  the  region  in  the  jcy-plane 
bounded  by  y  =  x  and  y  =  x2  —  x. 

37.  Find  the  volume  under  the  plane  z  =  4x  +  2y  +  25 
and  over  the  region  in  the  xy-plane  bounded  by  y  = 
x2  -  10  and  y  =  31  -  (x  -  \  f. 

38.  (a)  Set  up  an  iterated  integral  to  compute  the  volume 

under  the  hyperbolic  paraboloid  z  =  x2  —  y2  +  5 
and  over  the  disk 


D  =  {(x,  y) 


x2  +  y2  <  4} 


in  the  xy -plane. 

(b)  Use  a  computer  algebra  system  to  evaluate  the 
integral. 

39.  Find  the  volume  of  the  region  under  the  graph  of 

f(x,y)  =  2-\x\-\y\ 
and  above  the  xy-plane. 

40.  (a)  Show  that  if  R  =  [a,  b]  x  [c,  d],  f  is  continuous 

on  [a,  b],  and  g  is  continuous  on  [c,  d],  then 


f(x)g(y)dA 


{f,md')  if.™") 


(b)  What  can  you  say  about 

f{x)g(y)dA 


11 


if  D  is  not  a  rectangle?  More  specifically,  what  if 
D  is  an  elementary  region  of  type  1  ? 


334       Chapter  5    Multiple  Integration 


41.  Let 


/(*.  y) 


1  if  x  is  rational 

0    if  x  is  irrational  and  y  <  1 

2  if  x  is  irrational  and  y  >  1 


(a)  Show  that  fQ  f(x,y)dy  does  not  depend  on 
whether  x  is  rational  or  irrational. 


(b)  Show  that 
value. 


fo  f(x>  y)dy  dx  exists  and  find  its 


(c)  Partition  R  =  [0,  1]  x  [0,  2]  and  construct  a 
Riemann  sum  by  choosing  "test  points"  Cy  in  each 
subrectangle  of  the  partition  to  have  rational  x- 


coordinates.  Then  to  what  value  must  this  Riemann 
sum  converge  as  both  Ax;  and  Ay  j  tend  to  zero? 

( d)  Partition  R  and  construct  a  Riemann  sum  by  choos- 
ing test  points  Cy  =  (xf,  y  *)  such  that  x*  is  rational 
if  y]  <  1  and  xf  is  irrational  if  y  *  >  1 .  What  hap- 
pens to  this  Riemann  sum  as  both  A.t,  and  At- 
tend to  zero? 

(e)  Show  that  /  fails  to  be  integrable  on  R  by  us- 
ing Definition  2.3.  Thus,  we  see  that  double  inte- 
grals and  iterated  integrals  are  actually  different 
notions. 


5.3   Changing  the  Order  of  Integration 

Frequently,  it  is  useful  to  think  about  the  evaluation  of  double  integrals  over 
elementary  regions  essentially  as  the  determination  of  an  appropriate  order  of 
integration.  When  the  region  of  integration  is  a  rectangle,  Fubini's  theorem 
(Theorem  2.6)  says  the  order  in  which  we  integrate  has  no  significance;  that  is, 

fffdA=f    f  f(x,y)dydx=f    f  f(x,y)dxdy. 

J  J  R  J a    J c  J c    J a 

(See  Figure  5.45.)  When  the  region  is  elementary  of  type  1  only,  we  must  integrate 
first  with  respect  to  y  (and  then  with  respect  to  x)  if  we  wish  to  evaluate  the  double 
integral  by  means  of  a  single  iterated  integral.  (See  Figure  5.46.)  Then 


fdA=       /  f(x,y)dydx 

J  JD  J  a  Jy(x) 


In  the  same  way,  when  the  region  is  elementary  of  type  2  only,  we  would  typically 
integrate  first  with  respect  to  x,  so  that 


r  r  rd  rP(y) 

fdA=        /  f(x,y)dxdy. 

J  JD  Jc  Ja(y) 


® 

® 

® 

® 

R  =  [a,  b]  x  [c,  d] 


x  =  a 

y  =  8(x) 
® 

® 

x  =  b 

y=y(x) 

Figure  5.45  Changing  the  order  of  integration  over  a 
rectangle. 


Figure  5.46  A  type  1  region 
leads  us  to  integrate  with  respect 
to  y  first. 


5.3  |  Changing  the  Order  of  Integration  335 


Figure  5.47  A  type  2  region 
leads  to  integration  with  respect 
to  x  first. 


Figure  5.48  The  region  D  of 
Example  1. 


Figure  5.49  Integrating  over 
the  region  D  of  Example  1  by 
integrating  first  with  respect  to  x. 


(See  Figure  5.47.)  When  the  region  is  elementary  of  type  3,  however,  we  can 
choose  either  order  of  integration,  at  least  in  principle.  Often,  this  flexibility  can 
be  used  to  advantage,  as  the  following  examples  illustrate: 

EXAMPLE  1  We  calculate  the  area  of  the  region  shown  in  Figure  5.48.  Con- 
sidering D  as  a  type  1  region,  we  obtain 


Area  of 


p  p  p  e    p  m  X 

D=         Id  A  (Why?)  =  /    /      1  dy  dx 

J  J D  J\  JO 

y  |  Q  x  dx  =  J   In  x  dx . 


The  single  definite  integral  that  results  gives  the  area  under  the  graph  of  y  =  In  x 
over  the  ^-interval  [1,  e],  just  as  it  should.  To  evaluate  this  integral,  we  need  to 
use  integration  by  parts:  Let  u  =  In  x  (so  du  =  1  /x  dx)  and  dv  =  dx  (so  v  =  x). 
Then 


Area  of  D 


[ 


In x  dx  =  In x  ■  x\ 


(VI 


dx 


(remember  f  u  dv  =  u  ■  v  —  f  v  du),  so 
Area  of  D  =  e  —  0 


j  dx  =  e  —  (e  —  1) 


1. 


Integration  by  parts  can  be  avoided  if  we  integrate  first  with  respect  to  x,  as 
schematically  suggested  by  Figure  5.49.  Hence, 


Area  of  D 


I  I  \dA  =  I  I  \dxdy=  I  x\eydy=  I  (e  —  ey)dy 
J  Jd  Jo  Je?  Jo     e  Jo 


=  (ey-e?)\l0  =  (e-e)-(0-e°)=l, 
which  checks  (just  as  it  should).  ♦ 

Note  that  the  two  iterated  integrals  we  used  to  calculate  the  area  in  Example  1 , 
namely, 

/» e   p  In  x  f^fe 
II      dy  dx    and     /     /  dxdy, 
J  i  Jo  Jo  Jey 

are  not  obtained  from  each  other  by  a  simple  exchange  of  the  limits  of  integration. 
The  only  time  such  an  exchange  is  justified  is  when  the  region  of  integration 
is  a  rectangle  of  the  form  [a,  b]  x  [c,  d]  so  that  all  limits  of  integration  are 
constants. 

EXAMPLE  2  Sometimes  changing  the  order  of  integration  can  make  an  im- 
possible calculation  possible.  Consider  the  evaluation  of  the  following  iterated 
integral: 


LI 


y  cos(x2)dx  dy. 


After  some  effort  (and  maybe  some  scratchwork),  you  should  find  it  impossible 
even  to  begin  this  calculation.  In  fact  it  can  be  shown  that  cos(x2)  does  not 
have  an  antiderivative  that  can  be  expressed  in  terms  of  elementary  functions. 
Consequently,  we  appear  to  be  stuck. 


Chapter  5  |  Multiple  Integration 


Figure  5.50  Note  that  x 


corresponds  to  y 
region  shown. 


fx  over  the 


On  the  other  hand,  it  is  easy  to  integrate  y  cos(x2)  with  respect  to  y.  This 
suggests  finding  a  way  to  change  the  order  of  integration.  We  do  so  in  two  steps: 

1.  Use  the  limits  of  integration  in  the  original  iterated  integral  to  identify  the 
region  D  in  R2  over  which  the  integration  takes  place.  (While  doing  this,  you 
should  make  a  wish  that  D  turns  out  to  be  a  type  3  region.) 

2.  Assuming  that  the  region  D  in  Step  1  is  of  type  3,  change  the  order  of 
integration. 

The  limits  of  integration  in  the  preceding  example  imply  that  D  can  be  described  as 

D  =  {(x,  y)  |  v2  <  x  <  4,  0  <  y  <  2}, 

as  suggested  by  Figure  5.50.  Now  Figure  5.50  can  be  used  to  change  the  order 
of  integration.  We  have 


a: 


ycos(x  )dxdy 


ff 

Jo  Jo 


y  cos(x  )dy  dx. 


It  is  now  possible  to  complete  the  calculation;  that  is, 


ff 

Jo  Jo 


y  cos(x2)  dy  dx 


-f 

-jfi 


—  cos(x2) 


cos(x  )  dx 


dx 


y=0 


16 


cos  u  du, 


where  u  =  x  and  du  =  2x  dx,  so  that,  finally, 

-2 


if 


ycos(x  )dxdy 


1    •       1 16 

4smMlo 


sin  16. 


The  technique  of  changing  the  order  of  integration  is  a  very  powerful  one,  but 
it  is  by  no  means  a  panacea  for  all  cumbersome  (or  impossible)  integrals.  It  relies 
on  an  appropriate  interaction  between  the  integrals  and  the  region  of  integration 
that  often  fails  to  occur  in  practice. 


5.3  Exercises 


1 .  Consider  the  integral 

r2x 


and  check  that  your  answer  agrees  with  part  (a). 

In  Exercises  2-9,  sketch  the  region  of  integration,  reverse  the 
order  of  integration,  and  evaluate  both  iterated  integrals. 


'-ff 

Jo  Jo 


(2  —  x  —  y  )  dy  dx 


j  j   {2x+\)dydx.  r2  r*-2* 

Jo  Jo 


y  dy  dx 


(a)  Evaluate  this  integral.  „2  m-/ 

(b)  Sketch  the  region  of  integration.  I    I       x  dx  dy 

(c)  Write  an  equivalent  iterated  integral  with  the  order 
of  integration  reversed.  Evaluate  this  new  integral  5.   /    /   (x  +  y)dxdy 
and  check  that  your  answer  agrees  with  part  (a).  Jo  Jjy 


i.  j  j  Idydx 


5.4  |  Triple  Integrals 


7.  J    J  exdxdy 

p7t/2  fCOSX 

i.    I       I       sinx  dydx 
Jo  Jo 


'  a: 


/4-y2 


y  dx  dy 


/4-v2 


When  you  reverse  the  order  of  integration  in  Exercises  10  and 
11,  you  should  obtain  a  sum  of  iterated  integrals.  Make  the 
reversals  and  evaluate. 

10.   f    [  (x-y)dvdx 
J -2  Jx2-2 


4  i-Ay-y1 


11 


u 


(y  +  \)dxdy 


In  Exercises  12  and  13,  rewrite  the  given  sum  of  iterated  in- 
tegrals as  a  single  iterated  integral  by  reversing  the  order  of 
integration,  and  evaluate. 


12 


r- 1    nx  p2  p2—x 

.   I    /sin  x  dy  dx  +  I     I      sin  x  dy  dx 
Jo  Jo  Ji  Jo 


13. 


/*/ 

Jo  Jo 


8    rJyft  yl2  i-^/y/i 

y  dx  dy  +  I      I        y  dx  dy 
Js  "'Vy=S 


In  Exercises  14—18,  evaluate  the  given  iterated  integral. 


•/'/ 

Jo  Jiy 
lo  ly 

■a 


cos  (x  )  dx  dy 


x  sin  xy  dx  dy 


smx 


dx  dy 


■  IT' 

Jo  Jo 

■  ff  • 

JO  Jy/2 


dy  dx 


dx  dy 


It  is  interesting  to  see  what  a  computer  algebra  system  does  with 
iterated  integrals  that  are  difficult  or  impossible  to  integrate  in 
the  order  given.  In  Exercises  19-21,  experiment  with  a  com- 
puter to  evaluate  the  given  integrals. 

19.  (a)  Determine  the  value  of  Jfl2  J^2  y2  cos  (xy)  dy  dx 

via  computer.  Note  how  long  the  computer  takes 
to  deliver  the  answer.  Does  the  computer  give  you 
a  useful  answer? 

(b)  If  you  were  to  calculate  the  iterated  integral  in  part 
(a)  by  hand,  in  the  order  it  is  written,  what  method 
of  integration  would  you  use?  (Don't  actually  carry 
out  the  evaluation,  just  think  about  how  you  would 
accomplish  it.) 

(c)  Now  reverse  the  order  of  integration  and  let  your 
computer  evaluate  this  iterated  integral.  Does  your 
computer  supply  the  answer  more  quickly  than  in 
part  (a)? 

20.  (a)  See     if     your     computer     can  calculate 

Jo  fx2  x  sm  (y2)  dy  dx  as  it  is  written, 
(b)  Now  reverse  the  order  of  integration  and  have 
your  computer  evaluate  your  new  iterated  integral. 
Which  of  the  computations  in  parts  (a)  or  (b)  is 
easier  for  your  computer? 

^^21.  (a)  Can  your  computer  evaluate /J  f^-i  ecosxdx  dy? 

(b)  Reverse  the  order  of  integration  and  have  it  try 
again.  What  happens? 


f—  y 


Figure  5.51  The  box 

B  =  [a,  b]  x  [c,  d]  x  [p,  q]. 


5.4  Triple  Integrals 

Let  f{x,  y,  z)  be  a  function  of  three  variables.  Analogous  to  the  double  integral, 
we  define  the  triple  integral  of  /  over  a  solid  region  in  space  to  be  the  limit  of 
appropriate  Riemann  sums.  We  begin  by  denning  this  integral  over  box-shaped 
regions  and  then  proceed  to  define  the  integral  over  more  general  solid  regions. 

The  Integral  over  a  Box  

Let  B  be  a  closed  box  in  R3  whose  faces  are  parallel  to  the  coordinate  planes. 
That  is, 

B  =  {{x,  y,  z)  eR3  |  a  <  x  <  b,  c  <  y  <d,  p  <  z  <q}. 
(See  Figure  5.51.)  We  also  use  the  following  shorthand  notation  for  B: 
B  =  [a,  b]  x  [c,  d]  x  [p,  q]. 


Chapter  5  |  Multiple  Integration 


b=x„ 


c  =  y0  y\  yi 
 y 


q  =  Zn 


P  =  Zo 


d  =  yn 


Figure  5.52  A  partitioned  box. 


DEFINITION  4.1  A  partition  of  B  of  order  n  consists  of  three  collections 
of  partition  points  that  break  up  B  into  a  union  of  n3  subboxes.  That  is,  for 
i,  j,  k  =  0, . . . ,  n,  we  introduce  the  collections  {jc,  },  {yj},  and  {zk},  such  that 

a  =  xo  <  x\  <  ■  ■  ■  <        <  Xj  <  ■  •  •  <  x„  =  b, 

c  =  yo  <  y\  <  ■  ■  ■  <  yj-i  <  yj  <  ■■■  <  yn  =  d, 

p  =  Zo  <  Z\  <  ■  ■  ■  <  Zk-l  <  Zk  <  ■  ■  ■  <  z„  =  q. 
(See  Figure  5.52.)  In  addition,  for  i,  j,k  =  1  n,  let 

Axj=Xi—Xi-u     Ayj  =  yj  —  yj-\,     and    Azk  =  Zk  -  Zk-u 


DEFINITION  4.2  Let  /  be  any  function  defined  on  B  =  [a,  b]  x  [c,  d]  x 
[p,  q].  Partition  B  in  some  way.  Let  Cy*  be  any  point  in  the  subbox 

Bijk  =  [xt-i,Xi]  x  [yj-uyj]  x  [z*-i.  Zifc]    (},j,k=  I,...,  n). 
Then  the  quantity 

n 

S=  f(cljk)AVijk, 

i,j,k=l 

where  AVfjk  =  AxjAyjAzk  is  the  volume  of  B^,  is  called  the  Riemann 
sum  of  j f  on  B  corresponding  to  the  partition. 


You  can  think  of  the  Riemann  sum  ^  f(Cijk)AVijk  as  a  weighted  sum  of 
volumes  of  subboxes  of  B,  the  weighting  given  by  the  value  of  the  function  /  at 
particular  "test  points"  Cy£  in  each  subbox. 


DEFINITION  4.3    The  triple  integral  of  /  on  £,  denoted  by 

///.'"• 

by    jjj  f(x,y,z)dV,    orby    jjj  f(x,y,z)dxdydz, 


5.4  |  Triple  Integrals 


/ 

o 

/ 

/ 

/ 

B 

Figure  5.53  The  subbox 
contributes  /(c,/a-)A V;,^ 
to  the  Riemann  sum  S.  If 
we  think  of  /  as 
representing  a  density 
function,  then  the  total 
mass  of  the  entire  box  B 
KffhfdV. 


is  the  limit  of  the  Riemann  sum  S  as  the  dimensions  Ax, ,  Ay, ,  and  Azk  of 
the  subboxes  Bjjk  all  approach  zero,  that  is, 

fff  fdV=         lim  T  /(c,;A)A  v,  Ay,  A-A. 

J  J  JB  allA«,A^,A«-.0,j^1 

provided  that  this  limit  exists.  When  fffBfdV  exists,  we  say  that  /  is 
integrable  on  B. 


The  key  point  to  remember  is  that  the  triple  integral  is  the  limit  of  Riemann 
sums.  It  is  this  notion  that  enables  useful  and  important  applications  of  integrals. 
For  example,  if  we  view  the  integrand  /  as  a  type  of  generalized  density  function 
("generalized"  because  we  allow  negative  density!),  then  the  Riemann  sum  S  is 
a  sum  of  approximate  masses  (densities  times  volumes)  of  subboxes  of  B.  These 
approximations  should  improve  as  the  subboxes  become  smaller  and  smaller. 
Hence,  we  can  use  the  triple  integral  fffg  f  dV,  when  it  exists,  to  compute  the 
total  mass  of  a  solid  box  B  whose  density  varies  according  to  /,  as  suggested  by 
Figure  5.53. 

Analogous  to  Theorem  2.5,  we  have  the  following  result  regarding  integra- 
bility  of  functions: 

THEOREM  4.4  If  /  is  bounded  on  B  and  the  set  of  discontinuities  of  /  on  B 
has  zero  volume,  then  fffBfdV  exists.  (See  Figure  5.54.) 


Figure  5.54  In 

Theorem  4.4  the 
discontinuities  of  /  on  B 
(shown  shaded)  must  have 
zero  volume. 


To  say  that  a  set  X  has  zero  volume  as  we  do  in  Theorem  4.4,  we  mean  that 

we  can  cover  X  with  boxes  B\,  B2  Bn, . . .  (i.e.,  so  that  X  c  [J™=1  Bn),  the 

sum  of  whose  volumes  can  be  made  arbitrarily  small. 

To  evaluate  a  triple  integral  over  a  box,  we  can  use  a  three-dimensional  version 
of  Fubini's  theorem. 


THEOREM  4.5  (Fubini's  theorem)  Let  /  be  bounded  on  B  =  [a,  b]  x 
[c,  d]  x  [p,  q]  and  assume  that  the  set  S  of  discontinuities  of  /  has  zero  vol- 
ume. If  every  line  parallel  to  the  coordinate  axes  meets  S  in  at  most  finitely  many 
points,  then 


///. 


fdV 


pi?    pel     pq  pb    pq  pel 

—  f{x,y,z)dzdydx=  III 

J  a    J  c    J  p  J  a    J  p    J  c 

pd    pb    pq  pd I    pq  pb 

=  /     /     /    f(x,y,z)dzdxdy=  /     /     /  f(x,y,z)dxdzdy 

J  c     J  a    J  p  J c    J  p    J a 


f(x,  y,  z)dy  dz  dx 


pq    pb    pd  pq    pd  pb 

=  I     I     I    f(x,y,z)dydxdz=  III  f(x,y,z)dxdydz. 

J  p    J  a    J  c  J  p    J  c    J  a 


q    pd  pb 


340       Chapter  5  |  Multiple  Integration 


Figure  5.55  The  function  /  is 
continuous  on  W. 


z=Y(x,y) 


z  =  (p(x,y) 


-y 


Figure  5.56  An  elementary 
region  of  type  1 . 


y  =  d 

x  =  a(y){^ 

y  =  c 

Figure  5.57 

The  "shadow" 

(projection)  of  W  into  the  .rj-plane 
should  be  an  elementary  region  in 
the  plane. 


EXAMPLE  1  Let 

B  =  [-2,  3]  x  [0,  1]  x  [0,  5],    and  let    f(x,  y,  z)  =  x2ey  +  xyz. 

Thus,  /  is  continuous  and  hence  certainly  satisfies  the  hypotheses  of  Fubini's 
theorem.  Therefore, 

f  f  [  (xV  +xyz)dV  =  f    f    f  (x2ey +xyz)dzdydx 

J  J  Jb  J -2  Jo  JO 

J    (x2eyz  +  \xyz2)\l=0  dy  dx 
=  j  j  (5x2ey  +  f  xy)  dy  dx 
=  ^  {5x2ey  +  2ixy%T^dx 

=  j   (5(e-  l)x2  +  fx)dx 

=  (f(e-iy  +  f^)|3_2 

=  (45(e-l)+f)-(-f(e-l)+f) 

=  f(,-D+f. 

You  can  check  that  integrating  in  any  of  the  other  five  possible  orders  produces 
the  same  result.  ♦ 


Elementary  Regions  in  Space  

Now  suppose  W  denotes  a  fairly  arbitrary  solid  region  in  space,  like  a  rock  or  a 
slab  of  tofu.  Suppose  /  is  a  continuous  function  defined  on  W,  such  as  a  mass 
density  function.  (See  Figure  5.55.)  Then  the  triple  integral  of  /  over  W  should 
give  the  total  mass  of  W.  As  was  the  case  with  general  double  integrals,  we  need  to 
find  a  way  to  properly  define  fffw  f  dV  and  to  calculate  it  in  practical  situations. 
The  course  of  action  is  much  like  before:  We  see  how  to  calculate  integrals  over 
certain  types  of  elementary  regions  and  treat  integrals  over  more  general  regions 
by  subdividing  them  into  regions  of  elementary  type. 


DEFINITION  4.6  We  say  that  W  is  an  elementary  region  in  space  if  it  can 
be  described  as  a  subset  of  R3  of  one  of  the  following  four  types: 

Type  1  (see  Figures  5.56  and  5.57) 

(a)  W  =  {(x,  y,  z)  |  <p(x,  y)  <  z  <  f(x,  y),  y(x)  <  y  <  S(x),  a  <x  <b], 

or 

(b)  W  =  {(x,  y,  z)  |  <p(x,  y)<z<  fix,  y),  a(y)  <x<  B(y),  c<y<d}. 


5.4  |  Triple  Integrals  341 


Type  2  (see  Figure  5.58) 

(a)  W  =  {(x,  y,  z)  |  a(y,  z)  <  x  <  P(y,  z),  y(z)  <y<  S(z),  p<z<q), 

or 

(b)  W  =  {(x,  y,  z)  |  a(y,  z)  <  x  <  p(y,  z),  <p(y)  <z<  f(y),  c<y<  d}. 
Type  3  (see  Figure  5.59) 

(a)  W  =  {(x,  y,  z)  |  y(x,  z)  <y  <  S(x,  z),  a(z)  <x<  fi(z),  p  <  z  <  q}, 

or 

(b)  W  =  {(x,  y,  z)  |  y{x,  z)<  y  <  S(x,  z),  (p(x)  <z<  ir(x),  a<x<b}. 


Type  4 


W  is  of  all  three  previously  described  types. 


x=  a(y,z) 


y=y(x,z) 


y  =  5(x, z) 


Figure  5.58  For  an  elementary 
region  of  type  2,  the  shadow  in  the 
jz-plane  should  be  an  elementary 
region  in  the  plane. 


Figure  5.59  For  an  elementary 
region  of  type  3,  the  shadow  in  the 
-tz-plane  should  be  an  elementary 
region  in  the  plane. 


Some  explanation  regarding  Definition  4.6  is  in  order.  An  elementary  region 
W  of  type  1  is  a  solid  shape  whose  top  and  bottom  boundary  surfaces  each  can 
be  described  with  equations  that  give  z  as  functions  of  x  and  y  and  such  that 
the  projection  of  W  into  the  xy-plane  (the  "shadow")  is  in  turn  an  elementary 
region  in  R2  (in  the  sense  of  Definition  2.8).  Similarly,  an  elementary  region  of 
type  2  is  one  whose  front  and  back  boundary  surfaces  each  can  be  described  with 
equations  giving  x  as  functions  of  y  and  z  and  whose  projection  into  the  yz-plane 
is  an  elementary  region  in  R2 .  Finally,  an  elementary  region  of  type  3  is  one  whose 
left  and  right  boundary  surfaces  each  can  be  described  with  equations  giving  y 
as  functions  of  x  and  z  and  whose  projection  into  the  xz-plane  is  an  elementary 
region  in  R2.  In  each  case,  an  elementary  region  in  space  is  one  for  which  we 
can  find  boundary  surfaces  described  by  equations  where  one  of  the  variables  is 
expressed  in  terms  of  the  other  two,  and  whose  "shadow"  in  the  plane  of  these 
two  variables  is  an  elementary  region  in  R2  in  the  sense  of  Definition  2.8. 

EXAMPLE  2  Let  W  be  the  solid  region  bounded  by  the  hemisphere  x2  +  y2  + 
z2  =  4,  where  z  <  0,  and  the  paraboloid  z  =  4  —  x2  —  y2.  (See  Figure  5.60.)  It 
is  an  elementary  region  of  type  1  since  we  may  describe  it  as 


W  =  {(x,  y,  z)  |         -x2-y2  <  z  <  4 

-V4  -  x2  <  y  <  ^4  -  x2,  -2  <  x  <  2j 


2  2 
x  -  y  , 


342       Chapter  5  |  Multiple  Integration 


z 


x2  +  y2  +  z2  =  4, 

z  <  0  Shadow  of  W 

Figure  5.60  The  solid  region  W  of  Example  2. 


This  description  was  obtained  by  noting  that  W  is  bounded  on  top  and  bottom  by  a 
pair  of  surfaces,  each  of  which  is  the  graph  of  a  function  of  the  form  z  =  g(x,  y) 
and  the  shadow  of  W  in  the  xy-plane  is  a  disk  D  of  radius  2,  which  we  have 
chosen  to  describe  as 


D=\(x,y)  |  -J*- 


<  y  < 


-2  <  x  <  2 


and  which  we  already  know  is  an  elementary  region  (of  type  3)  in  the  xy- 
plane.  ♦ 

EXAMPLE  3   The  solid  bounded  by  the  ellipsoid 

a,b,c  positive  constants 


2  2  2 

^      XL        V  Z 

E  :  —  +  J—  +  — 


b2 


1. 


can  be  seen  to  be  an  elementary  region  of  type  4.  To  see  that  it  is  of  type  1 ,  split 
the  boundary  surface  in  half  via  the  z  =  0  plane  as  shown  in  Figure  5.61.  (This 
is  accomplished  analytically  by  solving  for  z  in  the  equation  for  the  ellipsoid.) 
Then  the  shadow  D  of  E  is  the  region  inside  the  ellipse  in  the  xy-plane  shown  in 
Figure  5.62. 


Figure  5.61  The  ellipsoid  of 
Example  3  as  an  elementary 
region  of  type  1 . 


Figure  5.62  The  shadow  of 
the  type  1  ellipsoid  in 
Figure  5.61  is  the  region  inside 
the  ellipse  x2/a2  +  y2/b2  =  1 
in  the  xy-plane. 


5.4  |  Triple  Integrals  343 


Figure  5.63  The  ellipsoid  of 
Example  3  as  a  type  2  elementary 
region. 


Figure  5.64  The  shadow 
of  the  ellipsoid  in 
Figure  5.63  is  the  region 
inside  the  ellipse 
y2/b2  +  z2/c2  =  1  in  the 
yz-plane. 


We  have 

D  =  \(x,  y) 


Figure  5.65  The  ellipsoid  of 
Example  3  as  a  type  3  region. 


-bJ\ 


y  <  bJi 


-a  <  x  <  a 


=    (*,  y) 


-a 


'l-^,  -b<  y<  b\ 


Figure  5.66  The 

shadow  of  the 
ellipsoid  in 
Figure  5.65  in  the 
Az-plane. 


so  D  is  in  fact  a  type  3  elementary  region  in  R2. 

To  see  that  E  is  of  type  2,  split  the  boundary  at  the  x  =  0  plane  as  in 
Figure  5.63.  The  shadow  in  the  yz-plane  is  again  the  region  inside  an  ellipse. 
(See  Figure  5.64.)  Finally,  to  see  that  E  is  of  type  3,  split  along  y  =  0.  (See 
Figures  5.65  and  5.66.)  ♦ 

Triple  Integrals  in  General  — 

Suppose  W  is  an  elementary  region  in  R3  and  /  is  a  continuous  function  on  W. 
Then,  just  as  in  the  case  of  double  integrals,  we  define  the  extension  of  /  by 


.f(x,y,z)  if(x,y,z)eW 
f   (x,  y,z)  =  { 

0  if(x,y,z)£W 


By  Theorem  4.4,  fext  is  integrable  on  any  box  B  that  contains  W.  Thus,  we  can 
make  the  following  definition: 


DEFINITION  4.7  Under  the  assumptions  that  W  is  an  elementary  region 
and  /  is  continuous  on  W,  we  define  the  triple  integral 

fff  fdV    tobe     fff  rxdV, 

J  J  Jw  J  J  J B 

where  B  is  any  box  containing  W. 


344       Chapter  5  |  Multiple  Integration 


Using  a  proof  analogous  to  that  of  Theorem  2.10,  we  can  establish  the 
following: 


THEOREM  4.8  Let  W  be  an  elementary  region  in  R3  and  /  a  continuous 
function  on  W. 


1.  If  W  is  of  type  1  (as  described  in  Definition  4.6),  then 


err  fb  rs(x)  rirQc>y) 

///   fdV=       /       /         f(x,y,z)dzdydx,  (type  la) 

J  J  JW  J  a    Jy(x)  J(fi(x,y) 


or 


r  r  r  rd  rP(y)  rf(x>y) 

ill  fdV=       /       /        f(x,y,z)dzdxdy.         (type  lb) 

J  J  JW  Jc     Ja(y)  J<p(x,y) 


2.  If  W  is  of  type  2,  then 


r  r  r  rl   r$(z)  rP(y,z) 

fdV=  /  f(x,y,z)dxdydz, 

J  J  JW  J p    Jy(z)  Ja(y,z.) 


or 


r  r  r  rd  rf(y)  rP(y,z) 

//  /   fdV=       /        /        f(x,y,z)dxdzdy.  (type  2b) 

J  J  JW  Jc    Jtp(y)  Ja(y,z.) 


3.  If  W  is  of  type  3,  then 


or 


r  r  r  ri   rP(z)  rHx,z) 

fdV=  /  f(x,y,z)dydxdz, 

J  J  JW  J  p    Ja(z)  Jy(x,z) 


err  rb  rfU)  rHx,z.) 

fdV=       /        /  f(x,y,z)dydzdx. 

J  J  JW  Ja    Jtp(x)  Jy(x,z) 


(type  2a) 


(type  3a) 


(type  3b) 


(0,0, 1). 


Plane  x  +  y  +  z  =  1 


.(0,1,0) 

-y 


(1,0,  0) 


Figure  5.67  The  tetrahedron  of 
Example  4. 


EXAMPLE  4   Let  W  denote  the  (solid)  tetrahedron  with  vertices  at  (0,  0,  0), 
(1,0,  0),  (0,  1,  0),  and  (0,  0,  1)  as  shown  in  Figure  5.67.  Suppose  that  the  mass 
density  at  a  point  (x,  y,  z)  inside  the  tetrahedron  varies  as  f(x,  y,  z)  —  1  +  xy. 
We  will  use  a  triple  integral  to  find  the  total  mass  of  the  tetrahedron. 
The  total  mass  M  is 


HI  fdV=  fff  (l+xy)dV. 
J  J  Jw  J  J  Jw 


(See  the  remark  before  Theorem  4.4.)  To  evaluate  this  triple  integral  using  iterated 
integrals,  note  that  we  can  view  the  tetrahedron  as  a  type  1  elementary  region. 
(Actually,  it  is  a  type  4  region,  but  that  will  not  matter.)  The  slanted  face  is  given 
by  the  equation  x  +  y  +  z  =  1,  which  describes  the  plane  that  contains  the  three 
points  (1,  0,  0),  (0,  1,  0),  and  (0,  0,  1).  Hence,  by  first  integrating  with  respect  to 


5.4  |  Triple  Integrals  345 


z  and  holding  x  and  y  constant, 


(0, 1,  0) , 

*  Line  x  +  y  =  1 
\^  (in  z  =  0  plane) 

(1,0,0) 

Figure  5.68  The  shadow  in  the 
xy-plane  of  the  tetrahedron  of 
Example  4  is  a  triangular  region. 

Figure  5.69  The  region  W  of 
Example  5. 


M=  f f       (f  \\+xy)dz\dA 

J  J  shadow  \J  0  / 

=  11  {l+xy){\-x-y)dA 

J  J  shadow 

=  /  /        (l  -  x  -  y  +  xy  -  x2y  -  xy2)  dA. 

J  J  shadow 


The  shadow  of  W  in  the  x  v-plane  is  just  the  triangular  region  shown  in  Figure  5.68. 
Thus, 


M  =  (l  -  x  —  y  +  xy  -  x2y  -  xy2)  dA 

J  J  shadow 

=  /     /       (l  —  x  —  y  +  xy  —  x2y  —  xy2)  dy  dx 
Jo  Jo 

=  f  ((1  -  je)  -  x(\  -  x)  -  1(1  -  x)2  +  \x{\  -  x  f 
Jo 


10 

L 


lx2(l  -  xf  -  \x{\-xf)dx 


=    G-f*  +  ^3-|*V*=(2 


1  _  5^  1 .3  _  lx4)dx  =  (lx  _  +  lx4  _   l_x5yL  _  1_ 


12' 


30      )\0  40" 


Note  that  M  can  also  be  written  as  a  single  iterated  integral,  namely, 


M 


pi    p  1— x    p  1 

Jo  Jo  Jo 


(1  +  xy)dz  dy  dx. 


EXAMPLE  5  We  calculate  the  volume  of  the  solid  W  sitting  in  the  first  octant 
and  bounded  by  the  coordinate  planes,  the  paraboloid  z  =  x2  +  y2  +  9,  and  the 
parabolic  cylinder  y  =  4  —  x2.  (See  Figure  5.69.) 

By  definition,  the  triple  integral  is  a  limit  of  a  weighted  sum  of  volumes  of 
tiny  subboxes  that  fill  out  the  region  of  integration.  If  the  weights  in  the  sum  are 
all  taken  to  be  1,  then  we  obtain  an  approximation  to  the  volume: 

i,j,k 


Therefore,  taking  the  limit  as  the  dimensions  of  the  subboxes  all  approach  zero, 
it  makes  sense  to  define 


v= ////""■ 


Chapter  5    Multiple  Integration 


(0,4) 


Figure  5.70  The  shadow 
in  the  xy-plane  of  the 
region  in  Figure  5.69. 


z  =  3x2  +  3y2-16 


gure  5.71  The  capsule-shaped 
gion  of  Example  6. 


Figure  5.72  The  shadow 
of  the  region  W  of 
Example  6,  obtained  by 
projecting  the  intersection 
curves  of  the  defining 
paraboloids  onto  the 
xy-plane. 


In  our  situation,  W  is  a  type  1  region  whose  shadow  in  the  xy-plane  looks 
like  the  region  shown  in  Figure  5.70.  Thus,  by  Theorem  4.4, 


V  =  dV  =       /        /  dzdydx 

J  J  Jw         Jo  Jo  Jo 

=  /    /       (x2  +  y2  +  9)dydx 
Jo  Jo 

=  f*(*ly  +  \ys  +  9y\^)dx 

=  f  (x2(4  -  x2)  +  i(4  -  x2)3  +  9(4  -  x2))  dx 
Jo 

=  f2  (m-2lx2  +  3x4  -  \x6)dx 


=  (ipx  _  7x3  +  fx 


J_v7)l  - 


21 " 


35 


EXAMPLE  6   We  find  the  volume  inside  the  capsule  bounded  by  the  paraboloids 
z  =  9  —  x2  —  y2  and  z  =  3x2  +  3y2  —  16.  (See  Figure  5.71.) 
Once  again,  we  have 


V 


IdV, 


and  again  the  region  W  of  interest  is  elementary  of  type  1 .  The  shadow,  or  pro- 
jection, of  W  in  the  xy-plane  is  determined  by 

{(x,  y)  e  R2  |  there  is  some  z  such  that  (x,  y,  z)  e  W}. 

Physically,  one  can  also  imagine  the  shadow  as  the  hole  produced  by  allowing  W 
to  "fall  through"  the  xy-plane.  In  other  words,  the  shadow  is  the  widest  part  of  W 
perpendicular  to  the  z-axis.  From  Figure  5.71,  one  can  see  that  it  is  determined 
by  the  intersection  of  the  two  boundary  paraboloids.  The  shadow  itself  is  shown 
in  Figure  5.72.  The  intersection  may  be  obtained  by  equating  the  z-coordinates 
of  the  boundary  paraboloids.  Therefore, 


9  -  x2  -  y2  =  3x2  +  3y2 


Thus,  by  Theorem  4.4, 


16 


4x2  +  4y2  =  25 

*2  +  y2  =  2{  =  (l)2- 


f  f  f  fS/2    r^/25/4-x2  p9-x2-y2 

V  =  dV  =         /    /  dzdydx 

J  J  Jw  J -5/2  J-Jli/A-x2  ^3jc2+3y2-16 


/•5/2    p^/25/4-x2  p9-x2-y2 

=  4  /       /  /  dz  dy  dx. 

Jo       Jo  J3x2+3v2 


v2-16 


This  last  iterated  integral  represents  the  volume  of  one  quarter  of  the  capsule. 
Hence,  we  multiply  its  value  by  4  to  obtain  the  total  volume.  The  reason  for  this 


5.4  I  Exercises  347 


manipulation  is  to  make  the  subsequent  calculations  somewhat  simpler  (although 
the  computation  that  follows  is  clearly  best  left  to  a  computer). 
We  compute 


V 


rS/2  *9-x*-f 

=  41      /  /  dzdydx 

J0       J0  Jix2+3y2-i6 


r5/2  p^25/4-x2 

=  4  /       /  (25 -4x2  -4y2)dydx 

Jo  Jo 


(f-x2)3/2)jx 


=  4^  U25-4x% 

/•5/2  / 


2^  /25  _  x2  _  4  ^25  _  v2\3/2 


25  _  r2  _  4  ^25 
4       A  3 


(f -x2)3/2)jx 
(f  -x2)3/2)jx 


=  4  (4-f)(f-x2)3^x 
Jo 

f5/2 

=  ¥/  (2i-x2f2dX. 
Jo 


Now  let  x  =  J  sin#,  so  dx  =  |  cos9  d9.  Then 

32  f^2  /5        \3  5  1250  f^2 


32  f^2  /5        V  5  1250  Z^2  4 

V  =  —  -cosS     -cos6d8  =  /      cos4  Ode 

3  Jo      \2        /   2  3  Jo 

rn/2  ( 

L  G 


1250  Z^2  /l  \2 
—  /      (-(1+COS20)  ^ 


625  /'7r/2  , 

 /      (1  +  2  cos  2(9  +  cos2  26)  d6 

6  Jo 


625 
~6~ 


625  r1* '  \ 

(6»  +  sin26»)|o/2  +  —  /  -(1  +  cos  46)  d6 

6  Jo  2 

12  (H  = 


625       625  (it      \  625n 

 7t  + 

12 


5.4  Exercises 


Evaluate  the  triple  integrals  given  in  Exercises  1—3. 


1.   f  [ [  xyzdV 

J  J  J[-l,l]x[0,2]x[l,3] 

2  SSL 


SSL 


(x2  +  y2  +  z2)dV 


3.  I  I   dV 

/[l,e]x[l,e]x[l,e] 

4.  Find  the  value  of  fffwzdV,  where  W  =  [—1,  2]  x 
[2, 5]  x  [—3, 3],  without  resorting  to  explicit 
calculation. 


/[0,l]x[0,2]x[0,3] 


348       Chapter  5  |  Multiple  Integration 


Evaluate  the  iterated  integrals  given  in  Exercises  5—7. 

5.  f    I    [  3yz2dxdydz 
J-i  Ji  Jo 

6.  J     I    I    (x  +  2y  +  z)  dy  dx  dz 

J\  Jo  Ji 

ply  ry+z 
JO   Jl+y  Jz 


z  dx  dz  dy 


8.  (a)  Let  W  be  an  elementary  region  in  R3.  Use  the 

definition  of  the  triple  integral  to  explain  why 
fffw  1  dV  gives  the  volume  of  W. 

(b)  Use  part  (a)  to  find  the  volume  of  the  region 
W  bounded  by  the  surfaces  z  =  x2  +  y2  and  z  = 
9-x2  -y2. 

9.  Use  triple  integrals  to  verify  that  the  volume  of  a  ball 
of  radius  a  is  4jra3/3. 

1 0.  Use  triple  integrals  to  calculate  the  volume  of  a  cone  of 
radius  r  and  height  It .  (You  may  wish  to  use  a  computer 
algebra  system  for  the  evaluation.) 

In  Exercises  11—20,  integrate  the  given  function  over  the  indi- 
cated region  W. 

11.  f(x,  y,  z)  —  2x  —  y  +  z;  W  is  the  region  bounded 
by  the  cylinder  z  =  y2,  the  Jty-plane,  and  the  planes 
x  =  0,  x  =  1,  y  =  —2,  y  =  2. 

12.  f(x,y,z)  =  y;  W  is  the  region  bounded  by  the 
plane  x  +  y  +  z  =  2,  the  cylinder  x2  +  z2  =  1,  and 
y  =  0. 

13.  f(x,y,z)  =  Sxyz;  W  is  the  region  bounded  by 
the  cylinder  y  =  x2,  the  plane  y  +  z  =  9,  and  the 
xy-plane. 

14.  f(x,y,z)  =  z",  W  is  the  region  in  the  first  octant 
bounded  by  the  cylinder  y2  +  z2  =  9  and  the  planes 
y  =  x,  x  =  0,  and  z  =  0. 

15.  f(x,  y,z)  =  1  —  z2',  W  is  the  tetrahedron  with  vertices 
(0,  0,  0),  (1,  0,  0),  (0,  2,  0),  and  (0,  0,  3). 

16.  f(x,  y,  z)  =  3.x:;  W  is  the  region  in  the  first  octant 
bounded  by  z  =  x2  +  y2,  x  =  0,  y  =  0,  and  z  =  4. 

17.  f(x,  y,  z)  =  x  +  y;  W  is  the  region  bounded  by 
the  cylinder  x2  +  3z2  =  9  and  the  planes  y  =  0, 
x  +  y  =  3. 

18.  f(x,  y,  z)  =  z;  W  is  the  region  bounded  by  z  =  0, 
x2  +  Ay2  =  4,  and  z  =  X  +  2. 

19.  f(x,  y,  z)  =  4jc  +  y;  W  is  the  region  bounded  by  x  = 

y2 ,  y  =  z,  x  =  y,  and  z  =  0. 

20.  f(x,  y,  z)  =  x;  W  is  the  region  in  the  first  octant 


21.  Find  the  volume  of  the  solid  bounded  by  z  =  4  —  x2, 
x  +  y  =  2,  and  the  coordinate  planes. 

22.  Find  the  volume  of  the  solid  bounded  by  the  planes  y  = 
0,  z  =  0,  2y  +  z  =  6,  and  the  cylinder  x2  +  y2  =  9. 

23.  Find  the  volume  of  the  solid  bounded  by  the  paraboloid 
z  =  Ax2  +  y2  and  the  cylinder  y2  +  z  =  2. 

24.  Find  the  volume  of  the  region  inside  both  of  the  cylin- 
ders x2  +  y2  =  a2  and  x2  +  z2  =  a2. 

25.  Consider  the  iterated  integral 

•  l-x 


lo 


f(x,  y,  z)dz  dx  dy. 


Sketch  the  region  of  integration  and  rewrite  the  inte- 
gral as  an  equivalent  iterated  integral  in  each  of  the  five 
other  orders  of  integration. 

26.  Change  the  order  of  integration  of 

f(x,  y,  z)dz dx  dy 


fff 

Jo  Jo  Jo 


to  give  five  other  equivalent  iterated  integrals. 
27.  Change  the  order  of  integration  of 


fff 

Jo  Jo  Jo 


f(x,  y,  z)dz  dy  dx 


bounded  by  z  =  x  +  2y  ,  z  =  6 
and  y  =  0. 


y  ,  x 


o, 


to  give  five  other  equivalent  iterated  integrals. 

28.  Consider  the  iterated  integral 

/    /  /  2dzdydx. 

Jo    Jo  Ay2 

(a)  This  integral  is  equal  to  a  triple  integral  over  a  solid 
region  W  in  R3.  Describe  W. 

(b)  Set  up  an  equivalent  iterated  integral  by  integrating 
first  with  respect  to  z,  then  with  respect  to  x,  then 
with  respect  to  y.  Do  not  evaluate  your  answer. 

(c)  Set  up  an  equivalent  iterated  integral  by  integrating 
first  with  respect  to  y,  then  with  respect  to  z,  then 
with  respect  to  x.  Do  not  evaluate  your  answer. 

(d)  Now  consider  integrating  first  with  respect  to  y, 
then  x,  then  z.  Set  up  a  sum  of  iterated  integrals 
that,  when  evaluated,  give  the  same  result.  Do  not 
evaluate  your  answer. 

(e)  Repeat  part  (d)  for  integration  first  with  respect  to 
x,  then  z,  then  y. 

29.  Consider  the  iterated  integral 

,2     n\sf4^  rA-y1 

/     /  /        (x3  +y3)dzdydx. 

3-2  Jo  Jx2+3y2 

(a)  This  integral  is  equal  to  a  triple  integral  over  a  solid 
region  W  in  R3.  Describe  W. 


5.5  I  Change  of  Variables  349 


(b)  Set  up  an  equivalent  iterated  integral  by  integrating 
first  with  respect  to  z,  then  with  respect  to  x,  then 
with  respect  to  y.  Do  not  evaluate  your  answer. 

(c)  Set  up  an  equivalent  iterated  integral  by  integrating 
first  with  respect  to  x,  then  with  respect  to  z,  then 
with  respect  to  y.  Do  not  evaluate  your  answer. 


(d)  Now  consider  integrating  first  with  respect  to  x, 
then  y,  then  z.  Set  up  a  sum  of  iterated  integrals 
that,  when  evaluated,  give  the  same  result.  Do  not 
evaluate  your  answer. 

(e)  Repeat  part  (d)  for  integration  first  with  respect  to 
y,  then  z,  then  x. 


5.5   Change  of  Variables 

As  some  of  the  examples  in  the  previous  sections  suggest,  the  evaluation  of  a  mul- 
tiple integral  by  means  of  iterated  integrals  can  be  a  complicated  process.  Both  the 
integrand  and  the  region  of  integration  can  contribute  computational  difficulties. 
Our  goal  for  this  section  is  to  see  ways  in  which  changes  in  coordinates  can  be 
used  to  transform  iterated  integrals  into  ones  that  are  relatively  straightforward  to 
calculate.  We  begin  by  studying  the  coordinate  transformations  themselves  and 
how  such  transformations  affect  the  relevant  integrals. 


Coordinate  Transformations   

Let  T:  R2  —>  R2  be  a  map  of  class  C1  that  transforms  the  wu-plane  into  the  xy- 
plane.  We  are  interested  particularly  in  how  certain  subsets  D*  of  the  wu-plane 
are  distorted  under  T  into  subsets  D  of  the  xy-plane.  (See  Figure  5.73.) 


D* 

CO 


D  =  T(D*) 


Figure  5.73  The  transformation  T(m,  v)  =  (x(u,  v),  y(u,  v)) 
takes  the  subset  D*  in  the  uv-p\am  to  the  subset  D  =  {(x,  y)  | 
(x,  y)  =  T(m,  v)  for  some  (u,  »)  e  D*)  of  the  xy-plane. 


EXAMPLE  1  Let  T(u,  v)  =  (u  +  1 ,  v  +  2);  that  is,  let  x  =  u  +  1,  y  =  v  +  2. 
This  transformation  translates  the  origin  in  the  Mu-plane  to  the  point  (1 ,  2)  in  the 
xy-plane  and  shifts  all  other  points  accordingly.  The  unit  square  D*  =  [0, 1]  x 
[0,  1  ],  for  example,  is  shifted  one  unit  to  the  right  and  two  units  up  but  is  otherwise 
unchanged  as  shown  in  Figure  5.74.  Thus,  the  image  of  D*  is  D  =  [1,2]  x  [2,3]. 

♦ 

EXAMPLE  2  Let  S(w,  v)  =  (2m,  3d).  The  origin  is  left  fixed,  but  S  stretches  all 
other  points  by  a  factor  of  two  in  the  horizontal  direction  and  by  a  factor  of  three 
in  the  vertical  direction.  (See  Figure  5.75.)  ♦ 

EXAMPLE  3    Composing  the  transformations  in  Examples  1  and  2,  we  obtain 

(T  o  S)(«,  «)  =  T(2«,  3u)  =  (2m  +  1,  3v  +  2). 
Such  a  transformation  must  both  stretch  and  translate  as  shown  in  Figure  5.76.  ♦ 


350       Chapter  5  |  Multiple  Integration 


D 


D 


□ 

=  T(D') 


Figure  5.74  The  image  of  D*  =  [0,  1]  x  [0,  1]  is 

D  =  [1,  2]  x  [2,  3]  under  the  translation 
T(m,  v)  =  (u  +  1,  v  +  2)  of  Example  1. 


£>*  =  [0, 1]  x  [0, 1] 


D  =  S(D*)  =  [0,  2]  x  [0,  3] 


Figure  5.75  The  transformation  S  of  Example  2  is  a  scaling 
by  a  factor  of  2  in  the  horizontal  direction  and  3  in  the  vertical 
direction. 


y 


ToS 


D 


D  =  [1,  3]  x  [2,  5] 


1 


Figure  5.76  Composition  of  the  transformations  of 
Examples  1  and  2. 


EXAMPLE  4  Let  T(w,  v)  =  (u  +  v,  u  —  v).  Because  each  of  the  component 
functions  of  T  involves  both  variables  u  and  v,  it  is  less  obvious  how  the  unit 
square  D*  =  [0,  1]  x  [0,  1]  transforms.  We  can  begin  to  get  some  idea  of  the 
geometry  by  seeing  how  T  maps  the  edges  of  D* : 


Bottom  edge:  (u,  0),  0  <  u  <  1 

Top  edge:  (u,  1),  0  <  u  <  1 

Left  edge:  (0,  v),  0  <  v  <  1 

Right  edge:  (1,  v),  0  <  v  <  1 


T(m,  0)  =  (u,  u); 
T(m,  1)  =  (u  +  1,  m  —  1); 
T(0,  v)  =  (v, -v); 
T(l,  w)  =  O  +  l,  1  -  v). 


By  sketching  the  images  of  the  edges,  it  is  now  plausible  that  the  image  of  D* 
under  T  is  as  shown  in  Figure  5.77.  ♦ 


Figure  5.77  The  transformation  T  of  Example  4. 


by 


5.5  |  Change  of  Variables  351 
More  generally,  we  consider  linear  transformations  T:  R2  — >  R2  defined 

T(w,  v)  =  (au  +  bv,  cu  +  dv) 


a  b 

u 

c  d 

V 

where  a,  b,  c,  and  d  are  constants.  (Note:  The  vector  (u,  v)  is  identified  with  the 
u 

v 


2  x  1  matrix 


.)  One  general  result  is  stated  in  the  following  proposition: 


PROPOSITION  5.1  Let  A  = 

fined  by 


a  b 
c  d 


,  where  det  A  ^  0.  If  T:  R2  ->  R2  is  de- 


T(w,  w)  =  A 


then  T  is  one-one,  onto,  takes  parallelograms  to  parallelograms  and  the  vertices 
of  parallelograms  to  vertices.  (See  §2. 1  to  review  the  notions  of  one-one  and  onto 
functions.)  Moreover,  if  D*  is  a  parallelogram  in  the  i/u-plane  that  is  mapped 
onto  the  parallelogram  D  =  T(D*)  in  the  xy-plane,  then 

Area  of  D  =  \  det  A|  ■  (Area  of  D*). 


EXAMPLE  5  We  may  write  the  transformation  T(u,  v)  =  (u  +  v,  u 
Example  4  as 


v)  in 


T(m,  v) 


Note  that 


det 


u 

v 


=  -2^0. 


Hence,  Proposition  5.1  tells  us  that  the  square  D*  =  [0,  1]  x  [0,  1]  must  be 
mapped  to  a  parallelogram  D  =  T(Z)*)  whose  vertices  are 

T(0,  0)  =  (0,  0),    T(0,  1)  =  (1,  -1),    T(1,0)  =  (1,1),    T(l,  1)  =  (2,0). 

Therefore,  Figure  5 .77  is  indeed  correct  and,  in  view  of  Proposition  5.1,  could  have 
been  arrived  at  quite  quickly.  Also  note  that  the  area  of  D  is  |  —  2|  •  1  =  2.  ♦ 

Proof  of  Proposition  5.1  First  we  show  that  T  is  one-one.  So  suppose 
T(w,  v)  =  T(u',  v').  We  show  that  then  u  =  u',v  =  v' .  We  have 

T(w,  v)  =  T(u' ,  v') 

if  and  only  if 

(au  +  bu ,  cu  +  dv)  =  (au'  +  bv',  cu'  +  du'). 
By  equating  components  and  manipulating,  we  see  this  is  equivalent  to  the  system 


a(u  —  u')  +  b(v  —  v')  =  0 
c(u  -  u')  +  d(v  -  v')  =  0 


(1) 


352       Chapter  5  |  Multiple  Integration 


If  a  7^  0,  then  we  may  use  the  first  equation  to  solve  for  u  —  u': 


u!  =  — (v-  v') 


(2) 


Figure  5.78  The  vertices  of 
D*  =  {p  +  sn  +  tb  |  0  <  s,t  < 
1 }  are  at  p,  p  +  a,  p  +  b, 
p  +  a  +  b  (i.e.,  where  s  and  t  take 
on  the  values  0  or  1). 


Figure  5.79  The  image  D  of  the 
parallelogram  D*  under  the  linear 
transformation  T(u)  =  An. 


Hence,  the  second  equation  in  (1)  becomes 

be 


(v  -  v')  +  d(v  -v')  =  0 


or,  equivalently, 


-be  +  ad 


(v  -  v')  =  0. 


By  hypothesis,  det  A  =  ad  —  be  ^  0.  Thus,  we  must  have  v  —  v'  =  0  and  there- 
fore, u  —  u'  =  0  by  equation  (2).  If  a  =  0,  then  we  must  have  both  b  0  and 
c     0,  since  det  A  ^  0.  Consequently,  the  system  (1)  becomes 

J  b(v  -v')  =0 
I  c(u  -  u')  +  d(v  -  v')  =  0  " 

The  first  equation  implies  v  —  v'  =  0  and  hence,  the  second  becomes  c(u  —  u')  = 
0,  which  in  turn  implies  u  —  u!  =  0,  as  desired. 

To  see  that  T  is  onto,  we  must  show  that,  given  any  point  (x,  y)  e  R2,  we  can 
find  (u,  v)  e  R2  such  that  T(m,  v)  =  (x,  y).  This  is  equivalent  to  solving  the  pair 
of  equations 

Iau  +  bv  =  x 
cu  +  dv  =  y 

for  u  and  v.  We  leave  it  to  you  to  check  that 

dx  —  by  ay  —  cx 

14  =  ~~a — TT    and    v  =  ~1 — TT 
ad  —  be  ad  —  be 

will  work. 

Now,  let  D*  be  a  parallelogram  in  the  ww-plane.  (See  Figure  5.78.)  Then  D* 
may  be  described  as 

D*  =  {u  |  u  =  p  +     +  tb,  0  <  s  <  1,  0  <  t  <  1}. 

Hence, 

D  =  T(D*)  =  {Au  ueD) 

=  {A(p  +  5a  +  rb)  |  0  <  s  <  1,  0<?<1} 

=  {Ap  +  ^Aa  +  rAb  |  0  <  s  <  1,  0  <  t  <  1}. 

If  we  let  p'  =  Ap,  a'  =  Aa,  and  b'  =  Ab,  then 

D  =  {p'  +  sa'  +  rb'  |  0  <  s  <  1,0  <  r  <  1}. 

Thus,  £>  is  also  a  parallelogram  and  moreover,  the  vertices  of  D  correspond  to 
those  of  D*.  (See  Figure  5.79.) 

Finally,  note  that  the  area  of  the  parallelogram  D*  whose  sides  are  parallel  to 


a2 


and  b 


by 

b2 


5.5  |  Change  of  Variables 


may  be  computed  as  follows: 


j 

k 

Areaof/?*  =  ||axb||  = 

det 

0 

bi 

b2 

0 

=  \a.\b2  —  a2b\ 


Similarly,  the  area  of  D  =  T(D*)  whose  sides  are  parallel  to 


a  = 


and   b'  = 


b\ 
b' 


is 


a'2b\  j 


a 

b 

a\ 

aa\  - 

-  ba2 

.  a2  . 

c 

d 

CL2 

ca\  - 

\-da2 

'  b\  ' 

a 

b  ~ 

"  bx  ' 

ab\  - 

Ybb2 

,b2_ 

c 

d 

.  b2  . 

cb\  - 

Ydb2 

Area  of  D  =  ||a'  x  b'||  =  \a[b'2 
Now,  a'  =  Aa  and  b'  =  Ab.  Therefore, 


and 


Hence,  by  appropriate  substitution  and  algebra, 

Area  of  D  =  \(aai  +  ba2)(cb\  +  db2)  —  (ca\  +  da-£)(ab\  +  bb2)\ 

=  \{ad  —  bc)(a\b2  —  a2b\)\ 

=  I  det  A]  •  area  of  D*. 

Note  that  we  have  not  precluded  the  possibility  of  D*'s  being  a  "degenerate" 
parallelogram,  that  is,  such  that  the  adjacent  sides  are  represented  by  vectors  a 
and  b,  where  b  is  a  scalar  multiple  of  a.  When  this  happens,  D  will  also  be 
a  degenerate  parallelogram.  The  assumption  that  det  A  7^  0  guarantees  that  a 
nondegenerate  parallelogram  D*  will  be  transformed  into  another  nondegenerate 
parallelogram,  although  we  have  not  proved  this  fact.  ■ 

Essentially  all  of  the  preceding  comments  can  be  adapted  to  the  three- 
dimensional  case.  We  omit  the  formalism  and,  instead,  briefly  discuss  an  example. 

EXAMPLE  6   Let  T:  R3  ->  R3  be  given  by 

T(m,  v,  w)  =  (2u,  2u  +  3v  +  w,  3w). 

Then  we  rewrite  T  by  using  matrix  multiplication: 


Note  that  if 


-  2 

0 

0  " 

u 

w) 

2 

3 

1 

V 

_  0 

0 

3  _ 

w 

"  2 

0 

0  " 

\  = 

2 

3 

1 

» 

0 

0 

3 

then  det  A  =  18^0. 


354       Chapter  5  |  Multiple  Integration 


A  result  analogous  to  Proposition  5.1  allows  us  to  conclude  that  T  is  one-one 
and  onto,  and  T  maps  parallelepipeds  to  parallelepipeds.  In  particular,  the  unit 
cube 

D*  =  [0,  1]  x  [0,  1]  x  [0,  1] 

is  mapped  onto  some  parallelepiped  D  =  T(£>*)  and,  moreover,  the  volume  of  D 
must  be 

|  det  A|  -  volume  of  D*  =  18-  1  =  18. 
To  determine  D,  we  need  only  determine  the  images  of  the  vertices  of  the  cube: 


r(o,  o,  0)  =  (o,  o,  o) 

T(0,  0,  1)  =  (0,  1,3) 
T(0,  1,  1)  =  (0,4,  3) 
Both  D*  and  its  image  D  are  shown  in  Figure  5.80. 


7X1,  0,  0)  =  (2,  2,  0);  T(0,  1,0)  =  (0,  3,  0); 
T(l,  1,  0)  =  (2,  5,  0);  7X1,  0,  1)  =  (2,  3,  3); 
J(l,  1,  1)  =  (2,  6,  3). 


D* 

Figure  5.80  The  cube  D*  and  its  image  D  under  the 
linear  transformation  of  Example  6. 

EXAMPLE  7   Of  course,  not  all  transformations  are  linear  ones.  Consider 

(x,  y)  =  T(r,  9)  =  (r  cos#,  r  sin#). 

Note  that  T  is  not  one-one  since  T(0,  0)  =  (0,  0)  =  T(0,  n).  (Indeed  T(0,  9)  = 
(0,  0)  for  all  real  numbers  9.)  Note  that  vertical  lines  in  the  r0-plane  given  by 
r  =  a,  where  a  is  constant,  are  mapped  to  the  points  (x,  y)  =  (a  cos#,  a  sin#) 
on  a  circle  of  radius  a.  Horizontal  rays  {(r,  9)  \  9  =  a,  r  >  0}  are  mapped  to 
rays  emanating  from  the  origin.  (See  Figure  5.81.)  It  follows  that  the  rectangle 
D*  =  [5,  1]  x  [0,  it]  in  the  r#-plane  is  mapped  not  to  a  parallelogram,  but  bent 


6 


G  =  a 

r  =  a 

Image  of 
6  =  a 


Image  of 
r  =  a 


Figure  5.81  The  images  of  lines  in  the  r#-plane  under  the 
transformation  T(r,  6)  =  (r  cos  0,  r  sin  0). 


5.5  |  Change  of  Variables  355 


D 


Figure  5.82  The  image  of  the  rectangle  D*  =  [±,  1]  x  [0,  n] 
under  T(r,  9)  =  (r  cos  6,  r  sin#). 


Figure  5.83  The  image  of  B*  =  [\,  1]  x  [0,  it]  x  [0,  1] 
under  T(r,  6,  z)  =  (r  cos  6,  r  sin#,  z). 


into  a  region  D  that  is  part  of  the  annular  region  between  circles  of  radii  \  and  1 , 
as  shown  in  Figure  5.82. 

Analogously,  the  transformation  T:  R3  — >  R3  given  by 

(x,  y,  z)  =  T(r,  6,  z)  =  (r  cos#,  r  sin^,  z) 

bends  the  solid  box  B*  =  [j,  1]  x  [0, 7r]  x  [0,  1]  into  a  horseshoe-shaped  solid. 
(See  Figure  5.83.)  ♦ 


Change  of  Variables  in  Definite  Integrals   

Now  we  see  what  effect  a  coordinate  transformation  can  have  on  integrals  and  how 
to  take  advantage  of  such  an  effect.  To  begin,  consider  a  case  with  which  you  are 
already  familiar,  namely,  the  method  of  substitution  in  single-variable  integrals. 

EXAMPLE  8  Consider  the  definite  integral  /Q2  2x  cos(x2)  dx.  To  evaluate,  one 
typically  makes  the  substitution  u  =  x2  (so  du  =  2x  dx).  Doing  so,  we  have 

sin  4. 


/  2x  cos(x2) dx  =  I  cosudu  =  sinu 
Jo  Jo 


u=4 
u=0 


Let's  dissect  this  example  more  carefully.  First  of  all,  the  substitution  u  =  x2 
may  be  rewritten  (restricting  x  to  nonnegative  values  only)  as  x  =  *Ju.  Then 
dx  =  du /(ly/u)  and 

f2  f4  du  f4 

/   2xcos(x2)dx=  I   2yfu  cos(a/m)2 — —  =  /   cosMJw  =  sin4. 
Jo  Jo  2y/u  Jo 

In  other  words,  the  substitution  is  such  that  the  2x  =  2^fu  factor  in  the  integrand 
is  canceled  by  the  functional  part  of  the  differential  dx  =  du/(2*Ju).  Hence,  a 
simple  integral  results.  ♦ 


In  general,  the  method  of  substitution  works  as  follows:  Given  a  (perhaps 
complicated)  definite  integral  fA  f(x)dx,  make  the  substitutions  =  x(u),  where 
x  is  of  class  C1.  Thus,  dx  =  x'(u)du.  If  A  =  x(a),  B  =  x(b),  and  x'(u)  7^  0  for 
u  between  a  and  b,  then 

/•  B  nb 

I    f(x)dx  =  J    f(x(u))x'(u)du.  (3) 

J  A  Ja 


Chapter  5  |  Multiple  Integration 


x 


x  =  x(u) 


It 


Figure  5.84  As  Am  =  du  — >•  0, 
Ax  — >  dx  =  x'(u)  Am.  Thus,  the  factor 
x'(u)  measures  how  length  in  the  w-direction 
relates  to  length  in  the  x-direction. 


Note  that  it  is  possible  to  have  a  >  b  in  (3)  above.  (This  happens  if  x(u)  is 
decreasing.)  Although  the  m -integral  in  equation  (3)  may  at  first  appear  to  be  more 
complicated  than  the  x -integral,  Example  8  shows  that  in  fact  just  the  opposite 
can  be  true. 

Beyond  the  algebraic  formalism  of  one-variable  substitution  in  equation  (3), 
it  is  worth  noting  that  the  term  x'(u)  represents  the  "infinitesimal  length  distortion 
factor"  involved  in  the  changing  from  measurement  in  u  to  measurement  in  x. 
(See  Figure  5.84.)  We  next  attempt  to  understand  how  these  ideas  may  be  adapted 
to  the  case  of  multiple  integrals. 

The  Change  of  Variables  Theorem  for  Double  Integrals 

Suppose  we  have  a  differentiable  coordinate  transformation  from  the  wu-plane  to 
the  xy-plane.  That  is, 

T:  R2  ->  R2,     T(w,  v)  =  (x(u,  v),  y(u,  v)). 


DEFINITION  5.2    The  Jacobian  of  the  transformation  T,  denoted 

d(x,  y) 

d(u,  v) ' 

is  the  determinant  of  the  derivative  matrix  DT(u, 

v).  That  is, 

dx  dx 

d(x'y)  =detDT(u,  u)  =  det 

d(u,  v) 

du  dv 
dy  dy 
du  dv 

dx  dy      dx  dy 
du  dv      dv  du 

The  notation  d(x,  y)/d(u,  v)  for  the  Jacobian  is  a  historical  convenience.  The 
Jacobian  is  not  a  partial  derivative,  but  rather  the  determinant  of  the  matrix  of 
partial  derivatives.  It  plays  the  role  of  an  "infinitesimal  area  distortion  factor" 
when  changing  variables  in  double  integrals,  as  in  the  following  key  result: 


5.5  |  Change  of  Variables 


y 


(f.f) 


(8,0) 

Figure  5.85  The  triangular  region 
D  of  Example  9. 


THEOREM  5.3  (Change  of  variables  in  double  integrals)  Let  D  and  D* 
be  elementary  regions  in  (respectively)  the  xy-plane  and  the  Mu-plane.  Suppose 
T:R2  — >  R2  is  a  coordinate  transformation  of  class  C1  that  maps  D*  onto  D 
in  a  one-one  fashion.  If  /:  D  — >  R  is  any  integrable  function  and  we  use  the 
transformation  T  to  make  the  substitution  x  =  x(u,  v),  y  =  y(u,  v),  then 


jj  f(x,y)dxdy  =  j j  f(x(u,v),y(u,v)) 


dp,  y) 
d(u,  v) 


du  dv. 


EXAMPLE  9   We  use  Theorem  5.3  to  calculate  the  integral 


cos(x  +  2y)  sin(x  —  y)  dx  dy 


over  the  triangular  region  D  bounded  by  the  lines  y  =  0,  y  =  x,andx  +  2y  =  8  as 
shown  in  Figure  5.85.  It  is  possible  to  evaluate  this  integral  by  using  the  relatively 
straightforward  methods  of  §5.2.  However,  this  would  prove  to  be  cumbersome, 
so,  instead  we  find  a  suitable  transformation  of  variables,  motivated  in  this  case  by 
the  nature  of  the  integrand.  In  particular,  we  let  u  =  x  +  2y,  v  =  x  —  y.  Solving 
for  x  and  y,  we  obtain 


u  +  2v 
x  =    and 


y 


Therefore, 


d(x,  y) 
d(u,  v) 


det 


Xu  Xy 

yu  yv 


det 


I  2 

3  3 

1  _  1 

3  3 


Considering  the  coordinate  transformation  as  a  mapping  T(w ,  v)  =  (x,  y)  of 
the  plane,  we  need  to  identify  a  region  D*  that  T  maps  in  a  one-one  fashion 
onto  D.  To  do  this,  essentially  all  we  need  do  is  to  consider  the  boundaries  of  D  : 


y  =  x 
x  +  2y  =  8 

y  =  0 


x  -  y  =  0 
u  =  8; 

u  —  v 
 =  0 


v  =  0; 


v  =  u. 


Hence,  one  can  see  that  T  transforms  the  region  D*  shown  in  Figure  5.86  onto 
D.  Therefore,  applying  Theorem  5.3, 


Figure  5.86  The  effect  of  the  transformation  T  of  Example  9. 


Chapter  5  |  Multiple  Integration 


j  j  cos(x  +  2  v)  sin(x  —  y)dx  dy  =  j j  cos  u  sin  v 


B(x,  y) 


du  dv 


d(u,  v) 

=  jj  cosw  sinv  |  — 1|  dudv 

=  J    J    j  cos  u  sin  v  dv  du 
Jo  Jo 

,8 

=  /    |  cos  u  ( —  cos  v)|  JJZq  du 
Jo 

=  |  /    cosw(—  cosm  +  1)  du 
Jo 

,  f* 

=  3/  (cos  u  —  cos  u)  du 
Jo 

sinwlg  —  j  j(1+cos2m)Jm 
sin 8  —  [ju  +  \  sin2M)|pJ 


=  \  [sin8  -  4 


sin 


16] 


There  is  another,  faster  way  to  calculate  the  Jacobian,  namely,  to  calculate 
d(u,  v)/d(x,  y)  directly  from  the  variable  transformation,  and  then  to  take  reci- 
procals. That  is,  from  the  equations  u  =  x  +  2y,  v  =  x  —  y,  we  have 


d(u,  v) 
d(x,  y) 


det 


I'v 


Uy 
Vy 


det 


-3. 


Consequently,  3(x,  y)/d(u,  v)  =  —  |,  which  checks  with  our  previous  result. 
This  method  works  because  if  T(h,  v)  =  (x,  v),  then,  under  the  assumptions  of 
Theorem  5.3,  (u,v)  =  T_1(jc,  y).  It  follows  from  the  chain  rule  that 


DT-\x,y)=  [DT(u,v)]-\ 

(That  is,  DT  1  is  the  inverse  matrix  of  DT.  See  Exercises  30-38  in  §  1 .6  for  more 
about  inverse  matrices.)  Hence, 


^=det[DT-1]=det[(DTr1] 

d(u,  v) 


detOT 


EXAMPLE  10  We  use  Theorem  5.3  to  evaluate  / fD(x2  -  y2)  exy  dx  dy,  where 
D  is  the  region  in  the  first  quadrant  bounded  by  the  hyperbolas  xy  =  1,  xy  =  4 
and  the  lines  y  =  x,  y  =  x  +  2.  (See  Figure  5.87.) 


5.5  |  Change  of  Variables  359 


y  y=x+2 


Figure  5.87  The  region  D  of  Example  10. 


D* 


Figure  5.88  The  region  D* 
corresponding  to  the  region  D  of 
Example  10. 


Both  the  integrand  and  the  region  present  complications  for  evaluation.  There 
would  seem  to  be  two  natural  choices  for  ways  to  transform  the  variables.  One 
would  be 

u  =  x2  —  y2    and    v  =  xy, 

motivated  by  the  nature  of  the  integrand.  However,  the  region  D  of  integration 
will  not  be  easy  to  describe  in  terms  of  this  particular  choice  of  wu-coordinates. 
Another  possible  transformation  of  variables,  motivated  instead  by  the  shape  of 
D,  is 

u  =  xy    and    v  =  y  —  x . 

Now  this  change  of  variables  would  not  seem  to  help  much  with  the  integrand, 
but,  as  we  shall  see,  it  turns  out  to  be  just  what  we  need. 

First  note  that  the  boundary  hyperbolas  xy  =  1  and  xy  =  4  correspond  re- 
spectively, to  the  lines  u  =  1  and  u  =  4;  the  lines  y  =  x  and  y  =  x  +  2  correspond 
to  v  =  0  and  v  =  2.  Thus,  the  region  D*  in  the  Mu-plane  that  corresponds  to  D 
(see  Figure  5.88)  is 

D*  =  {(u,  v)  |  1  <  u  <  4,  0  <  v  <  2}. 
Next,  we  calculate  that  the  Jacobian  of  the  variable  transformation  is 


9(w,  u) 


=  det 


y  x 
-1  1 


=  x  +  y. 


d(x,  y) 

Hence,  the  Jacobian  we  require  in  order  to  use  Theorem  5.3  is 

d(x,  y)  =  1 
9(w,  v)      x  +  y 

Moreover,  since  we  will  be  working  in  the  first  quadrant  (where  x  and  y  are  both 
positive),  |3(jc,  y)/9(w,  v)\  =  l/(x  +  y). 


Chapter  5  |  Multiple  Integration 


At  last  we  are  ready  to  compute: 


jji*2-  y2)  e*y  dx  dy  =  jj  (x2  -  y2)  e*y  (v^)  du  dv 
=  f2  C  (*-?)(*  +  y) 

Jo  J  i  x 


+  y 

du  dv 


exy  dudv 


-u(e4  -  el)dv  =  -y  0  -  e  ) 


2 

k' 


2(e  -  A 


Note  that  the  insertion  of  the  Jacobian  in  the  integrand  caused  precisely  the  can- 
celation needed  to  make  the  evaluation  straightforward.  We  cannot  always  expect 
this  to  happen,  but  the  lesson  here  is  to  be  willing  to  carry  through  calculations 
that  may  not  at  first  appear  to  be  so  easy.  ♦ 

EXAMPLE  11  (Double  integrals  in  polar  coordinates)  In  Example  9,  a  coor- 
dinate transformation  was  chosen  primarily  to  simplify  the  integrand  of  the  double 
integral.  In  this  example  we  change  variables  by  using  a  coordinate  system  better 
suited  to  the  geometry  of  the  region  of  integration. 

For  example,  suppose  that  the  region  D  is  a  disk  of  radius  a : 

D={(x,y)\x2  +  y2  <  a} 

=  |(x>  y)  I  —  V a2  —  x2  <  y  <  \l a1  —  x2,  —a  <  x  <  a  J  . 

Then,  to  integrate  any  (integrable)  function  /  over  D  in  Cartesian  coordinates, 
one  would  write 

/  /  f(x,y)dxdy=  I     I  f(x,y)dydx. 

J  Jd  J -a  J-Ja^l? 

Even  if  it  is  easy  initially  to  find  a  partial  antiderivative  of  the  integrand,  the  limits 
in  the  preceding  double  integral  may  complicate  matters  considerably.  This  is  be- 
cause the  disk  is  described  rather  awkwardly  by  Cartesian  coordinates.  We  know, 
however,  that  it  has  a  much  more  convenient  description  in  polar  coordinates  as 

{(r,  0)  |  0  <r  <a,0  <6  <  lit). 

This  suggests  that  we  make  the  change  of  variables 

(x,  y)  =  T(r,  0)  =  (r  cos#,  r  sin#), 

which  is  shown  in  Figure  5.89.  (Note  that  T  maps  all  points  of  the  form  (0,  6)  to 
the  origin  in  the  xy-plane  and  thus,  cannot  map  D*  in  a  one-one  fashion  onto  D. 
Nonetheless,  the  points  of  D*  on  which  T  fails  to  be  one-one  fill  out  a  portion 
of  a  line — a  one-dimensional  locus — and  it  turns  out  that  it  will  not  affect  the 
double  integral  transformation.)  The  Jacobian  for  this  change  of  variables  is 


^  =  det 
d(r,  6) 


cos  9  —rsinO 
sin  0     r  cos  9 


r  cos2  9  +  r  sin2  9  =  r. 


2k 

D* 

a 

5.5  |  Change  of  Variables 

y 


361 


-~^x2  +  y2 

A 

) 

Figure  5.89  T  maps  the  (nonclosed)  rectangle  D*  to 
the  disk  D  of  radius  a. 


< 

Figure  5.90  The  disk  of 
radius  a  centered  at  the 
origin. 


(0,2) 


x2  +  y2  =  4 


Figure  5.91  The  region 
D  of  Example  13. 


(Note  that  r  >  0  on  D,  so  \r\  —  r.)  Thus,  using  Theorem  5.3,  the  double  integral 
can  be  evaluated  by  using  polar  coordinates  as  follows: 


y)  dx  dy 


/a     p  \J a2—x2 
/   f(x,y)dydx 
-a  J—fa^1 


p2n  pa 
JO  JO 


f(r  cos#,  r  sin9)r  dr  d9. 


It  is  evident  that  the  limits  of  integration  of  the  r0-integrals  are  substantially 
simpler  than  those  in  the  xy-integral.  Of  course,  the  substitution  in  the  integrand 
may  result  in  a  more  complicated  expression,  but  in  many  situations  this  will  not  be 
the  case.  Polar  coordinate  transformations  will  prove  to  be  especially  convenient 
when  dealing  with  regions  whose  boundaries  are  parts  of  circles.  ♦ 

EXAMPLE  12  To  see  polar  coordinates  "in  action,"  we  calculate  the  area  of  a 
circle,  using  double  integrals.  Once  more,  let  D  be  the  disk  of  radius  a,  centered 
at  the  origin  as  in  Figure  5.90.  Then  we  have 


Area 


r-   r-  faf  V«2—  X2  pin  pa 

ldA=       /  dydx  = 

J  JD  J -a  J-Ja2-x2  JO  Jo 


r  dr  d9, 


following  the  discussion  in  Example  1 1 .  The  last  iterated  integral  is  readily  eval- 
uated as 


j    jrdrd9  =  j     (jr2\tyd9  =       \a2  dd  =  \a2(2Tt  -  0)  =  ita2 , 

which  indeed  agrees  with  what  we  already  know.  If  you  feel  so  inclined 
compare  this  calculation  with  the  evaluation  of  the  iterated  integral  in  Cartesian 
coordinates.  No  doubt  you'll  agree  that  the  use  of  polar  coordinates  offers  clear 
advantages.  ♦ 


EXAMPLE  13   We  evaluate  the  double  integral  ffD    x2  +  y2  +  ldx  dy, 

where  D  is  the  quarter  disk  shown  in  Figure  5.91,  using  polar  coordinates.  The 
region  D  of  integration  is  given  in  Cartesian  coordinates  by 


so  that 


D  =  {(x,  y)  |  0  <  y  <  y^-x2,  0  <  x  <  2}, 


f  f  yjx2  +  y2  +  \dxdy  =  f  f  y 1  x2  +  y2  +  \dydx 
J  Jd  Jo  Jo 


Chapter  5  |  Multiple  Integration 


This  iterated  integral  is  extremely  difficult  to  evaluate.  However,  D  corresponds 
to  the  polar  region 

D*  =  {(r,  9)\0<r  <2,  0  <  6>  <  jt/2}. 
Therefore,  using  Theorem  5.3,  we  have 

J  J  Vjc2  +  y2  +  1  dx  dy  =  ff^> 


r2  cos2  9  +  r2  sin2  9  +  1  •  r  dr  d6 


r2+\rdrd9 


=  r '  n, 

Jo  Jo 

=  /  |(r2+l)3/2|_0^ 
Jo 

-L 


i(53/2  -  \)dd 
-  !)■ 


Sketch  of  a  proof  of  Theorem  5.3  Let  (uq,  vq)  be  any  point  in  D*  and  let 
Au  =  u  —  uo,  Av  =  v  —  vo-  The  coordinate  transformation  T  maps  the  rectangle 
R*  inside  D*  (shown  in  Figure  5.92)  onto  the  region  R  inside  D  in  the  xy-plane. 
(In  general,  R  will  not  be  a  rectangle.)  Since  T  is  of  class  C1 ,  the  differentiability 
of  T  (see  Definition  3.8  of  Chapter  2)  implies  that  the  linear  approximation 


h(u,  v)  =  T(w0,  v0)  +  DT(m0,  Wo) 
=  T(u0,  vq)  +  DT(u0,  u0) 


U  —  Mo 

v  -  V0 


Au 
Av 


is  a  good  approximation  to  T  near  the  point  («o ,  fo)-  In  particular,  h  takes  the  rect- 
angle R*  onto  some  parallelogram  P  that  approximates  R  as  shown  in  Figure  5.93. 
We  compare  the  area  of  R*  to  that  of  P. 

From  Figure  5.93,  we  see  that  the  rectangle  R*  is  spanned  by 

0 


Am 
0 


and  b 


5.5  |  Change  of  Variables 


P  =  h(R*) 


DT(u0,  v0)b 


R  =  T(R') 


T(w0, 

h(Mo,  v0)        ^DT(u0,  u0)a 


Figure  5.93  The  linear  approximation  h  takes  the  rectangle  R*  onto  a  parallelogram  P 
that  approximates  R  =  T(R*). 


and  the  parallelogram  P  is  spanned  by  the  vectors  c  =  DT(mo,  t>o)a  and  d  = 
DT(wo,  wo)b.  Hence, 


Area  of  7?*  =  ||a  x  b||  =  Am  Av, 
and  thus,  by  Proposition  5.1, 

d(x,  y) 


Area  of  P  —  lie  x  dll 


detDT(w0,  Vo)\Au  Av 


d(u,  v) 


(u0,  v0) 


Au  Av. 


This  result  gives  us  some  idea  how  the  Jacobian  factor  arises. 

To  complete  the  sketch  of  the  proof,  we  need  a  partitioning  argument.  Partition 
D*  by  subrectangles  Rf, .  Then  we  obtain  a  corresponding  partition  of  D  into  (not 
necessarily  rectangular)  subregions  Rjj  =  T(7?*  ).  Let  A  Ajj  denote  the  area  of  Rjj . 
Let  Cij  denote  the  lower  left  corner  of  /?*■  and  let  di;  =  T(c,7).  (See  Figure  5.94.) 
Then,  since  /  is  integrable  on  D, 

/  /  f(x,y)dxdy=    lim  V/(d,7)AAi7. 

J  JD  all  Rij-tO  *7-f 


From  the  remarks  in  the  preceding  paragraph,  we  know  that 

d(x,y) 


AAjj  «a  area  of  parallelogram  h(/?* )  = 


d(u,  v) 


(ci7) 


Aui  Avj. 


Figure  5.94  A  partition  of  D*  gives  rise  to  a  partition  of  D. 


364       Chapter  5  |  Multiple  Integration 


Figure  5.95  The  "area  element" 
d A  in  rectangular  coordinates  is 
dx  dy. 


D' 


x  = r  cos  6 
y  =  r  sin  0 


Figure  5.96  The  polar-rectangular  transformation  takes  rectangles  in  the 
r#-plane  to  wedges  of  disks  in  the  Jty-plane. 


Arclength  =  rAd 


Figure  5.97  An  infinitesimal 
polar  wedge. 


Taking  limits  as  all  the  Rjj  tend  to  zero  (i.e.,  as  Am,  and  Avj  approach  zero),  we 
find  that 


[  [  f(x,y)dxdy  =     hm     Tf  (T(c,7)) 


d(x,  y) 


d(u,  v) 


(«y) 


All:  AVj 


If. 


f(x(u,  v),  y(u,  u)) 


d(x,  y) 


d(u,  v) 


du  dv, 


as  was  to  be  shown. 


Consider  again  the  polar-rectangular  coordinate  transformation.  When  we  use 
Cartesian  (rectangular)  coordinates  to  calculate  a  double  integral  over  a  region 
D  in  the  plane,  then  we  are  subdividing  D  into  "infinitesimal"  rectangles  having 
"area"  equal  to  dx  dy.  (See  Figure  5.95.)  On  the  other  hand  when  we  use  polar 
coordinates  to  describe  this  same  region,  we  are  subdividing  D  into  infinitesimal 
pieces  of  disks  instead.  (See  Figure  5.96.)  These  disk  wedges  arise  from  trans- 
formed rectangles  in  the  r#-plane.  One  such  infinitesimal  wedge  in  the  xy-plane 
is  suggested  by  Figure  5.97.  When  AG  and  Ar  are  very  small,  the  shape  is  nearly 
rectangular  with  approximate  area  (r  AO)  Ar.  Thus,  in  the  limit,  we  frequently  say 


d  A  =  dx  dy     (Cartesian  area  element) 
=  r  dr  d6    (polar  area  element). 


Change  of  Variables  in  Triple  Integrals   

It  is  not  difficult  to  adapt  the  previous  reasoning  to  the  case  of  triple  integrals.  We 
omit  the  details,  stating  only  the  main  results  instead. 


DEFINITION  5.4  Let  T:  R3  ->  R3  be  a  differentiable  coordinate  transfor- 
mation 

T(m,  v,  w)  =  (x(u,  v,  w),  y(u,  v,  w),  z(u,  v,  w)) 


5.5  |  Change  of  Variables 


from  www-space  to  xvz-space.  The  Jacobian  of  T,  denoted 

90,  y,  z) 


d(u,  v,  w)' 


is  det(DT(w,  v,  w)).  That  is, 


d(x,  y,  z) 


d(u,  v,  w) 


=  det 


3x 

dx 

dx 

du 

dv 

dw 

3? 

dy 

dy_ 

du 

dv 

dw 

dz 

dz 

dz 

_  du 

dv 

dw  _ 

In  general,  given  any  differentiate  coordinate  transformation  T:  R" 
the  Jacobian  is  just  the  determinant  of  the  derivative  matrix: 


R" 


d(xu  ...,x„) 


d{u\ 


detDT(M! 


3xi 

3xi 

3xi 

du\ 

du2 

du„ 

dxi 

3x2 

dx2 

det 

du\ 

du2 

dun 

dxn 

dx„ 

dxn 

du\ 

dU2 

dun 

THEOREM  5.5  (Change  of  variables  in  triple  integrals)   Let  W  and  W*  be 

elementary  regions  in  (respectively)  xvz-space  and  wuw-space,  and  let  T:  R3  -> 
R3  be  a  coordinate  transformation  of  class  C1  that  maps  W*  onto  W  in  a  one-one 
fashion.  If  /:  W  — >  R  is  integrable  and  we  use  the  transformation  T  to  make  the 
substitution  x  =  x(u,  v,  w),  y  =  y(u,  v,  w),  z  =  z(u,  v,  w),  then 


/  /  /  f{x,y,z)dxdydz 
J  J  Jw 


I  I Iw* 


f(x(u,  v,  w),  y(u,  v,  w),  z(u,  v,  w)) 


d(x,  y,  z) 
d(u,  v,  w) 


du  dv  dw. 


(See  Figure  5.98.) 


In  the  integral  formula  of  the  change  of  variables  theorem  (Theorem  5.5),  the 
Jacobian  represents  the  "volume  distortion  factor"  that  occurs  when  the  three- 
dimensional  region  W  is  subdivided  into  pieces  that  are  transformed  boxes  in 
Muw-space.  (See  Figure  5.99.)  In  other  words,  the  differential  volume  elements 


Chapter  5  |  Multiple  Integration 


Figure  5.98  A  three-dimensional  transformation  T  that  takes  the  solid 
region  W*  in  uvw-space  to  the  region  W  in  xyz-space. 


(i.e.,  "infinitesimal"  pieces  of  volumes)  in  xyz-  and  Muw-coordinates  are  related 
by  the  formula 


dV  =  dx  dy  dz  = 


d(x,  y,  z) 
d(u,  v,  w) 


dudvdw. 


Figure  5.99  The  volume  of  the  "infinitesimal  box"  in 
uvw-space  is  du  dv  dw.  The  image  of  this  box  under  T 
has  volume  \d(x,  y,  z)/d(u,  v,  w)\  du  dv  dw. 


EXAMPLE  14  (Triple  integrals  in  cylindrical  coordinates)  When  integrat- 
ing over  solid  objects  possessing  an  axis  of  rotational  symmetry,  cylindrical 
coordinates  can  be  especially  helpful.  The  cylindrical-rectangular  coordinate 
transformation 

'  x  =  r  cos  9 
y  =  r  sin  6 

z  =  z 


has  Jacobian 


d(x,  y,  z) 


=  det 


cost' 
sinO 
0 


-r  sind 
r  cos  9 
0 


=  r  cos  9  +  r  sin  9  =  r. 


d(r,  9,  z) 

Hence,  the  formula  in  Theorem  5.5  becomes 

f(r  cos6»,  r  sm.9,  z)r  dr  d9  dz. 


Ill   fix,  y,  z)dxdydz  = 
J  J  Jw  J  J  JV\ 


5.5  |  Change  of  Variables  367 

In  particular,  we  see  that  the  volume  element  in  cylindrical  coordinates  is 

dV  =  rdrd6dz. 

(Recall  that  the  cylindrical  coordinate  r  is  usually  taken  to  be  nonnegative.  Given 
this  convention,  we  may  omit  the  absolute  value  sign  in  the  change  of  variables 
formula.)  The  geometry  behind  this  volume  element  is  quite  plausible:  A  "dif- 
ferential box"  in  r#z-space  is  transformed  to  a  portion  of  a  solid  cylinder  that  is 
nearly  a  box  itself.  (See  Figure  5. 100.)  ♦ 


Figure  5.1 00  A  "differential  box"  in  rOz-space  is  mapped  to  a  portion  of  a 
solid  cylinder  in  xyz-space  by  the  cylindrical-rectangular  transformation. 


EXAMPLE  15  To  calculate  the  volume  of  a  cone  of  height  h  and  radius  a,  we 
may  use  Cartesian  coordinates,  in  which  case  the  cone  is  the  solid  W  bounded  by 
the  surface  az  =  h^/x2  +  y2  and  the  plane  z  =  h,  as  shown  in  Figure  5.101.  The 
volume  can  be  found  by  calculating  the  iterated  triple  integral 


/a  p  ~J a1—  x2  p 
-a  J—Ja^x2  J% 


dz  dy  dx. 


We  will  forgo  the  details  of  the  evaluation,  noting  only  that  trigonometric  substi- 
tutions are  necessary  and  that  they  make  the  resulting  computation  quite  tedious. 


az 


Shadow  in  xy-plane 


Figure  5.101  The  solid  cone  W  of  Example  15. 


In  contrast,  since  the  cone  has  an  axis  of  rotational  symmetry,  the  use  of 
cylindrical  coordinates  should  afford  us  substantially  less  involved  calculations. 


Chapter  5  |  Multiple  Integration 


Figure  5.102  The  cone  of 
Example  15  described  in 
cylindrical  coordinates. 


Hence,  we  consider  the  cone  again.  (See  Figure  5.102.)  Note  that 

h 


W  = 


(r,  9,  z) 


r  <  z  <h,  0  <  r  <a,  0  <  9  <  lit 


Thus,  the  volume  is  given  by 


n    n    n  p  7,71     no  nh 

dV  =         /     /  rdzdrdO. 

J  J  Jw  J0      J0  J±r 


(Note  the  order  of  integration  that  we  chose.)  The  evaluation  of  this  iterated 
integral  is  exceedingly  straightforward;  we  have 


/      /    /    rdzdrd0=  /  / 

Jo      Jo    3\r  Jo  Jo 

-L 

-re- 


2tt    Pa      /  h 

r\h-  -r\drd6 


h  j  h 

—r  1 

2  3a 


dO 


a  xd6 


=  2tc  I  —a' 
6 


=  —a  h, 


which  agrees  with  what  we  already  know. 


EXAMPLE  16  (Triple  integrals  in  spherical  coordinates)  Ifa  solid  object  has 
a  center  of  symmetry,  then  spherical  coordinates  can  make  integration  over  such 
an  object  more  convenient.  The  spherical-rectangular  coordinate  transformation 

x  =  p  sin  <p  cos  6 
y  =  p  sin  cp  sin  9 
z  =  p  cos  cp 


has  Jacobian 

d(x,  y,  z) 

B(p,<p,e) 


=  det 


simp  cost 
sin<p  sin£ 
coscp 


p  coscp  cos( 
p  cos  cp  sin  ( 
— p  sin^ 


-p  sin<p  sin# 
p  sin  <p  cos  0 
0 


Using  cofactor  expansion  about  the  last  row,  this  determinant  is  equal  to 
cos  (p  (p2  cos2  9  sin  <p  cos  cp  +  p2  sin2  9  sin  cp  cos  <p) 


+  p  sin  cp  (p  cos2  9  sin2  (p  +  p  sin2 1 


sin 


=  p2  cos  <p(sin  (p  cos  cp)  +  p2  sin3  <p 
=  p2  sin  cp  (cos2  cp  +  sin2  (p) 
=  p2  sincp. 

(Under  the  restriction  that  0  <  <p  <  jt,  sin^)  will  always  be  nonnegative.  Hence, 
the  Jacobian  will  also  be  nonnegative.)  Therefore,  the  volume  element  in  spherical 


5.5  |  Change  of  Variables 


de 


dq> 


/dp 


rdd  =  psin(p  dd 


Figure  5.1 03  A  differential  box  in  p<p#-space  is  mapped  to  a  portion  of  a  solid  ball 
in  xyz-space  by  the  spherical-rectangular  transformation. 


coordinates  is 


dV  =  p  sin<p  dp  d<p  dO, 


and  the  change  of  variables  formula  in  Theorem  5.5  becomes 

/  /  /   /(*>  y>  z)dxdydz 
J  J  Jw 


-III 


f(x(p,  ip,  0),  y(p,  <p,  0),  z(p,  (p,  0))p  sirup  dp  dip  dO. 


The  volume  element  in  spherical  coordinates  makes  sense  geometrically,  because 
a  differential  box  in  p(p0 -space  is  transformed  to  a  portion  of  a  solid  ball  that  is 
approximated  by  a  box  having  volume  p2  sin  q>  dp  d(p  dO.  (See  Figure  5. 103.)  ♦ 

EXAMPLE  17  The  volume  of  a  ball  is  easy  to  calculate  in  spherical  coordi- 
nates. A  solid  ball  of  radius  a  may  be  described  as 

B  =  {(p,  (p,9)\0<p<a,0<(p<7T,0<e<  2jt}. 

(See  Figure  5 . 1 04.)  Hence,  we  may  compute  the  volume  by  using  the  triple  integral 


fffdV  =  f!'  r  r 

J  J  Jb  Jo    Jo  Jo 


p  sirup  dp  d(p  d9 


,  .  3 
^0  ^0 


sin<p  dcp  dO 


-  (-coscp\*)d0  =  -  (-(-l)  +  l)d0 
j  Jo  j  Jo 


2a 


3  flit 


dO 


as  expected. 


EXAMPLE  18  We  return  to  the  example  of  the  cone  of  radius  a  and  height  h 
and  this  time,  use  spherical  coordinates  to  calculate  its  volume.  First,  note  two 
things :  (i)  that  the  cone 's  lateral  surface  has  the  equation  (p  =  tan~ 1  (a  /  h)  in  spher- 
ical coordinates  and  (ii)  that  the  planar  top  having  Cartesian  equation  z  =  h  has 
spherical  equation  p  cos  cp  =  h  or,  equivalently,  p  =  h  sec  (p.  (See  Figure  5. 105.) 


Chapter  5  |  Multiple  Integration 


For  fixed  values  of  the  spherical  angles  tp  and  0,  the  values  of  p  that  give 
points  inside  the  cone  vary  from  0  to  h  sec  (p.  Any  points  inside  the  cone  must 
have  spherical  angle  <p  between  0  and  tan~'(a/  h).  Finally,  by  symmetry,  0  can 
assume  any  value  between  0  and  2n ,  Hence,  the  cone  may  be  described  as  the  set 


0  <  p  <  h  sec<p,  0  <  (p  <  tan  1  '-,  0  <  0  <  2n  \ 

h 


Therefore,  we  calculate  the  volume  as 

"2jt    /»tan    (a/H)  phsecy 


pin    /'tan    \ajn)  p 

Jo     Jo  Jo 

=/7 

^0  ^0 


p  sin  <p  dp  d<p  d0 
/h)  (h  sec<p)3 


sin<pd(p  dO 


^3     r2n  ntarTl(a/h) 


hi     pm  pi 

=  —  /      /  secJ  cp  siri(p d(p dO 

3  Jo  Jo 

^3     pin  ptsarl(alh) 

=  —  /      /  tmcp  sec2  cp  dtp  dO. 

3  Jo  Jo 

Now,  let  u  =  tany>  so  du  =  sec2  (p  dep.  Then  the  last  integral  becomes 

h3  f2n  [a,h  h3  f2n  1  /a\2         h3a2  f2lt 

—  /  ududO  =  —  -  -  )  dO  =  — -  /  dO 
3  Jo    Jo  3  Jo    2  \hj  6h*  Jo 


3 

a2h 


-{lit) 


Tt 


-a  h, 


as  expected. 


The  use  of  spherical  coordinates  in  Example  18  is  not  the  most  appropri- 
ate. We  merely  include  the  example  so  that  you  can  develop  some  facility  with 
"thinking  spherically."  Further  practice  can  be  obtained  by  considering  some  of 
the  applications  in  the  next  section  as  well  as,  of  course,  some  of  the  exercises. 

Summary:  Change  of  Variables  Formulas   


Change  of  variables  in  double  integrals: 

jj  f(x,y)dxdy  =  j  j  f(x(u,v),y(u,v 


)) 


d(x,  y) 


d(u,  v) 


du  dv 


Area  elements: 


dA  =  dx  dy 
=  rdrd0 

djpc,  y) 
d(u,  v) 


(Cartesian) 
(polar) 

dudv  (general) 


5.5  I  Exercises  371 


Change  of  variables  in  triple  integrals: 

/  /  /   /(*>  J>  z)dxdydz 
J  J  Jw 


=fll 


f(x(u,  v,  w),  y(u,  v,  w),  z(u,  v,  w)) 


d(x,  y,  z) 


d(u,  v,  w) 


du  dvdw 


Volume  elements: 

dV  =  dxdydz  (Cartesian) 
=  r  dr  d6  dz  (cylindrical) 
=  p2  sin  (p  dp  d<p  d9  (spherical) 
3(x,  y,  z) 


d(u,  v,  w) 


dudvdw  (general) 


5.5  Exercises 


1.  LetT(M,  v)  =  (3m, -v). 
(a)  Write  T(u,  v)  as  A 


for  a  suitable  matrix  A. 


(b)  Describe  the  image  D  =  T(D*),  where  D*  is  the 
unit  square  [0,  1]  x  [0,  1]. 


2.  (a)  Let 


T(m,  v) 


u  —  v  u  +  v 


V5  '  V2 


How  does  T  transform  the  unit  square  D* 
[0,  1]  x  [0,  1]? 

(b)  Now  suppose 


T(k,  v) 


U  +  V    u  —  V 


V2  '  V2 
Describe  how  T  transforms  D*. 


3.  If 


T(m,  v)  = 


2  3 
-1  1 


and  D*  is  the  parallelogram  whose  vertices  are  (0,  0), 
(1,  3),  (-1,  2),  and  (0,  5),  determine  D  =  T(D*). 

4.  If  D*  is  the  parallelogram  whose  vertices  are  (0,  0), 
(—1,3),  (1,  2),  and  (0,  5)  and  D  is  the  parallelogram 
whose  vertices  are  (0,  0),  (3,  2),  (1,  —1),  and  (4,  1), 
find  a  transformation  T  such  that  T(Z)*)  =  D. 

5.  If  T(m,  v,  w)  =  (3m  —  v,  u  —  v  +  2w,  5m  +  3d  —  w), 
describe  how  T  transforms  the  unit  cube  W*  = 
[0,  1]  x  [0,  1]  x  [0,  1]. 


6.  Suppose  T(m,  d)  =  (u,  uv).  Explain  (perhaps  by  us- 
ing pictures)  how  T  transforms  the  unit  square  D*  = 
[0,  1]  x  [0,  1].  Is  T  one-one  on  D*l 

7.  Let  T:  R3  — >■  R3  be  the  transformation  given  by 

T(p,  <p,  6)  =  (p  sin  <p  cos  6,  p  sin<p  sin 9,  p  cos^). 

(a)  Determine   D  =  T(D*),  where   D*  =  [0,  1]  x 
[O.tt]  x  [0,2jt]. 

(b)  Determine   D  =  T(D*),  where   D*  =  [0,  1]  x 
[0, 7T/2]  x  [0,  tt/2]. 

(c)  Determine  D  =  T(D*),  where  D*  =  [1/2,  1]  x 
[0,  tt/2]  x  [0,  TT/2]. 

8.  This  problem  concerns  the  iterated  integral 


J0  Jy/2 


(2x  —  y)dx  dy. 


(a)  Evaluate  this  integral  and  sketch  the  region  D  of 
integration  in  the  .v  v-plane. 

(b)  Let  u  =  2x  —  y  and  v  =  y.  Find  the  region  D*  in 
the  MD-plane  that  corresponds  to  D. 

(c)  Use  the  change  of  variables  theorem  (Theorem  5.3) 
to  evaluate  the  integral  by  using  the  substitution 

u  =  2x  —  y,  v  =  y. 


9.  Evaluate  the  integral 

r2  |-(x/2)+l 

x/2 

by  making  the  substitution  u  =  x,  v  =  2y  —  x. 


ff 

Jo  Jxl 


x5(2y  -x)e(2y~x)2  dydx 


Chapter  5  |  Multiple  Integration 


10.  Determine  the  value  of 


x  +  y 


2y 


dA, 


where  D  is  the  region  in  R2  enclosed  by  the  lines 
y  =  x/2,  y  =  0,  and  x  +  y  =  1 . 

1 1 .  Evaluate  f  fD(2x  +  y)2ex~y  dA,  where  D  is  the  region 
enclosed  by  2x  +y  =  1,  2x  +  y  =  4,  x  —  y  =  —  1, 
and  x  —  y  =  1 . 


12.  Evaluate 


SL 


(2x  +  y-3)2 

  dx  dy, 

D  (2y  -x  +  6)2 


where  D  is  the  square  with  vertices  (0,  0),  (2,  1), 
(3,  -1),  and  (1,  -2).  (Hint:  First  sketch  D  and  find 
the  equations  of  its  sides.) 

In  Exercises  13—1 7,  transform  the  given  integral  in  Cartesian 
coordinates  to  one  in  polar  coordinates  and  evaluate  the  polar 
integral. 


3  dy  dx 


dy  dx 


Jo  Jo 

15.  J  j  (x2 +  y2)3/2dA,  where  D  is  the  disk*2 +  v2  <9 

/a  r^fa1 
-a  JO 

f'f 

Jo  Jo 


dx  dy 


dy  dx 
Jx2  +  y2 


18.  Evaluate 


11 


d  ^4" 


dA, 


where  D  is  the  disk  of  radius  1  with  center  at  (0,  1). 
(Be  careful  when  you  describe  D.) 

19.  Let  D  be  the  region  between  the  square  with  vertices 
(1,  1),(-1,  1),(-1,  -1),(1,  -1)  and  the  unit  disk  cen- 


tered at  the  origin.  Evaluate 


y2  dA. 


20.  Find  the  total  area  enclosed  inside  the  rose  r  =  sin  26 . 
(Hint:  Sketch  the  curve  and  find  the  area  inside  a  single 
leaf.) 

21.  Let  n  be  a  positive  integer,  and  let  a  be  a  posi- 
tive constant.  Calculate  the  total  area  inside  the  rose 
r  =  a  cosn8  and  show  that  the  value  depends  only  on 
a  and  whether  n  is  even  or  odd. 


22.  Find  the  area  of  the  region  inside  both  of  the  circles 
r  =  2a  cos  6  and  r  =  2a  sin  9,  where  a  is  a  positive 
constant. 

23.  Find  the  area  of  the  region  inside  the  cardioid  r  = 
1  —  cos  8  and  outside  the  circle  r  =  1 . 

24.  Find  the  area  of  the  region  bounded  by  the  positive 
x-axis  and  the  spiral  r  =  39,  0  <  6  <  2n. 


25.  Evaluate 


cos(x2  +  y2)dA, 


where  D  is  the  shaded  region  in  Figure  5.106. 


Arc  of  a  circle 
of  radius  1 
(centered  at 
origin) 


Figure  5.106  The  region  D  of 
Exercise  25. 


26.  Evaluate  j  j  sin  (x2  +  y2)dA,  where  D  is  the  region 

in  the  first  quadrant  bounded  by  the  coordinate  axes 
and  the  circles  x2  +  y2  =  1  and  x2  +  y2  =  9. 


27.  Use  polar  coordinates  to  evaluate 


where  D  is  the  unit  square  [0,  1]  x  [0,  1]. 

*3    /.V9-JC2  z-3 


dA, 


/)     fV-r  pi 
I  j 
-3  J  -^9=3?  J \s/x2+y1 

ina 

/l    r-s/\-y2  r4 
-1  J-Jl-v2  JO 


using  cylindrical  coordinates. 


Jx2  +  y 


dz  dy  dx  by 


29.  Evaluate 


using  cylindrical  coordinates. 
30.  Evaluate 

dV 


■  dz  dx  dy  by 


///. 


Ib  y/x2  +  y2  +  z2  +  3 ' 
where  B  is  the  ball  of  radius  2  centered  at  the  origin. 
31.  Determine 

(x2  +  y2+2z2)dV, 


where  W  is  the  solid  cylinder  defined  by  the  inequali- 
ties x2  +  y2  <  4,  - 1  <  z  <  2. 


5.6  |  Applications  of  Integration 


32.  Determine  the  value  of 


SSL 


dV ',  where 


>w  jx2  +  y2 

W  is  the  solid  region  bounded  by  the  plane  z  =  12  and 
the  paraboloid  z  =  2x2  +  2y2  —  6. 

33.  Find  the  volume  of  the  region  W  bounded  on  top  by 
z  =  y/a2  —  x2  —  y2,  on  the  bottom  by  the  xy-plane, 
and  on  the  sides  by  the  cylinder  x2  +  y2  =  b2,  where 

0  <  b  <  a. 

In  Exercises  34  and  35,  determine  the  values  of  the  given  in- 
tegrals, where  W  is  the  region  bounded  by  the  two  spheres 
x2  +  y2  +  z2  =  a2  and  x2  +  y2  +  z2  =  b2,  for  0  <  a  <  b. 


«•  SSL 


dV 


35 


w  jx2  +  y2  +  z2 

■III* 


x1  +  yL  +  zz  ex 


36.  Let  W  denote  the  solid  region  in  the  first  octant  be- 
tween the  spheres  x2  +  y2  +  z2  =  a2  and  x2  +  y2  + 
z2  =  b2,  where  0  <  a  <  b.  Determine  the  value  of 

fffw(x  +  y  +  z)dv. 

37.  Determine  the  value  of  fffw  z2  dV,  where  W  is  the 
solid  region  lying  above  the  cone  z  =  ^3x2  +  3y2  and 
inside  the  sphere  x2  +  y2  +  z2  =  6z. 


38.  Determine 


where  W 


(l  +  Jx2  +  y2)  dV, 


(x,y,z)\  y/x2  +  y2  <z/2<3\. 


39.  Find  the  volume  of  the  region  W  that  represents  the 
intersection  of  the  solid  cylinder  x  +  y2  <  1  and  the 
solid  ellipsoid  2(x2  +  y2)  +  z2  <  10. 

40.  Find  the  volume  of  the  solid  W  that  is  bounded  by 
the  paraboloid  z  =  9  —  x2  —  y2 ,  the  xy-plane,  and  the 
cylinder  x2  +  y2  =  4. 


41.  Find 


SSL 


(2  +  x2  +  y2)dV, 


where  W  is  the  region  inside  the  spheres2  +  y2  +  z2  = 
25  and  above  the  plane  z  =  3. 

42.  Find  the  volume  of  the  intersection  of  the  three  solid 
cylinders 

x2  +  y2<a2,     x2  +  z2<a2,     and  y2+z2<a2. 

(Hint:  First  draw  a  careful  sketch,  then  note  that,  by 
symmetry,  it  suffices  to  calculate  the  volume  of  a  por- 
tion of  the  intersection. ) 


Day 

°F 

Monday 

65 

Tuesday 

63 

Wednesday 

52 

Thursday 

51 

Friday 

45 

Saturday 

43 

Sunday 

47 

5.6  Applications  of  Integration 

In  this  section,  we  explore  a  variety  of  settings  where  double  and  triple  integrals 
arise  naturally. 

Average  Value  of  a  Function 

Suppose  temperatures  (shown  in  the  adjacent  table)  are  recorded  in  Oberlin,  Ohio, 
during  a  particular  week.  From  these  data,  we  calculate  the  average  (or  mean) 
temperature: 

65  +  63  +  52  +  51  +45  +  43  +  47 


Average  temperature 


52.3°F. 


Of  course,  this  calculation  only  represents  an  approximation  of  the  true  average 
value,  since  the  temperature  will  vary  during  each  day.  To  determine  the  true 
average  temperature,  we  need  to  know  the  temperature  as  a  function  of  time  for 
all  instants  of  time  during  that  one-week  period;  that  is,  we  consider 

Temperature  =  T(x),    x  =  elapsed  time  (in  days),    for    0  <  x  <  7. 

Then  a  more  accurate  determination  of  the  average  temperature  is  as  an  integral: 


Average  temperature 


=  1  fn* 

Jo 


)  dx. 


(1) 


Since  an  integral  is  nothing  more  than  the  limit  of  a  sum,  it's  not  hard  to  see  that 
the  preceding  formula  is  a  generalization  of  the  original  discrete  sum  calculation 
to  the  continuous  case.  (See  Figure  5.107.) 


374       Chapter  5  |  Multiple  Integration 


80  T 


10-- 


0        1        2        3        4        5  6 
Days 

Figure  5.1 07  A  continuous  temperature  function  T(x) 
over  the  interval  [0,  7].  The  average  temperature  for  the 
week  is  ^  /Q7  T(x)dx. 


Note  that 


7  =  /   dx  =  length  of  time  interval. 
Jo 

Hence,  we  may  rewrite  formula  (1)  as 

Jq  T(x)dx 


Average  temperature  = 


fo  dx 


This  observation  leads  us  to  make  the  following  definitions  concerning  average 
values  of  functions. 


DEFINITION  6.1  (a)  Let  /:  [a,  b]^Rbe  an  integrable  function  of  one 
variable.  The  average  (mean)  value  of  /  on  [a,  b]  is 


i"  ft   W     -  faf^dX  _  £  /W* 

U\«*-u     ,1    J  Wdx-  "length  of  interval  [a,  i]- 

J  a 


-f 

-a  J a 


(b)  Let  /:  D  c  R2  ->  R  be  an  integrable  function  of  two  variables.  The 
average  value  of  /  on  D  is 

m  _IfDfdA_ffDfdA 

ffDdA   ~  areaof/3' 

(c)  Let  /:  W  c  R3  — >  R  be  an  integrable  function  of  three  variables.  The 
average  value  of  /  on  W  is 

m        IffwfdV  ffUfdV 

L/Javg 


JffwdV       volume  of  W 


EXAMPLE  1  Suppose  that  the  "temperature  function"  for  Oberlin  during  a 
week  in  April  is 

T(r\  —  JIlv7  -  Mr6  -I-  1127  r5  _  2393    4   ,    66821    3  _  45781    2   ,    12581  , 

J  W  —  5040A         180"*  180  A  72  *    "1"    720  "*  360  A    "1"    210  *  "1"  UJ' 


5.6  |  Applications  of  Integration 


(0,1) 


(2,0) 


Figure  5.108  The  triangular 
metal  plate  of  Example  2. 


where  0  <  x  <  7.  Then  the  mean  temperature  for  that  week  would  be 


ITJavg  —  710  /  (5040*7 
JO 


107  6  1  1127  5 
180  180 


2393  4 
72  X 


_i_  66821  3 
720  X 

1  (  113  „8 


45781  „2   ,  12581 


360 


210 


x  +  65)  c/x 


40320 

,    66821  4 
2880  A 

888709 
17280 


107  7  1  1127  6 
1260  1080 


2393  5 
360 


^  +  ^  +  65x)\l 


1080 

51.43°F 


EXAMPLE  2  Suppose  that  the  thickness  of  the  triangular  metal  plate,  shown 
in  Figure  5.108,  varies  as  f(x,  y)  =  xy  +  1,  where  (x,  y)  are  the  coordinates  of 
a  point  in  the  plate.  The  average  thickness  of  the  plate  is,  therefore, 

Jo  fo~2y(xy  +  l)dxdy 


Average  thickness  = 


fofo  2y  dxdy 


Note  that 


1  r2-2y 


II 

JO  JO 


dx  dy  =  area  of  triangular  plate  =  ~(2  •  1)  =  1 


from  elementary  geometry.  Hence,  the  average  thickness  is 


fo  fo  ( 


xy+l)dxd\         I      i\    7  \|-*=2-2v  , 

1        =  /  (2*  y+x)\x=o  dy 

Jo 

=  f  (\{2-2yfy  +  (2-2y))dy 
Jo 

=  2j\y3-2y2+l)dy  =  2^- §  y3  +  y) 


EXAMPLE  3  (See  also  Example  6  of  §5.4.)  Suppose  the  temperature  inside  the 
capsule  bounded  by  the  paraboloids  z  =  9  —  x2  —  y2  and  z  =  3x2  +  3y2  —  16 
varies  from  point  to  point  as 

T(x,y,z)  =  z(x2  +  y2). 

We  calculate  the  mean  temperature  of  the  capsule. 
From  Definition  6.1, 


[-^lavg  — 


WwTdV 
IffwdV  ' 


The  particular  iterated  integrals  we  can  use  for  the  computation  are  then 

pS/2    ^25/4-x2  ^-x2-y2 

/       /  /  z(x2  +  y2)dzdydx 

J -5/2  J-^/25/4-A-2  J3x2+3y2-l6 

[-^]avg  = 


/5/2    p  y/25/4-x2  p9-x2-y2 
I  I  dzdy  dx 

-5/2  J-^/25/4-x2  Av2+3j-2-16 


Chapter  5  |  Multiple  Integration 


Unfortunately,  the  calculations  involved  in  evaluating  these  integrals  are  rather 
tedious. 

On  the  other  hand,  since  the  capsule  has  an  axis  of  rotational  symmetry, 
cylindrical  coordinates  can  be  used  to  simplify  the  computations.  Note  that  the 
boundary  paraboloids  have  cylindrical  equations  of  z  =  9  —  r2  and  z  =  3r2  —  16 
and  that  the  shadow  of  the  capsule  in  the  z  =  0  plane  can  be  described  in  polar 
coordinates  as 

{(r,0)  |  0  <  r  <  §,0  <  9  <  2n)  . 
(See  Figures  5.109  and  5.1 10.) 


{(r,6)\0<r<j,Q<e<2n} 

y 


z  =  3x2  +  3.y2-16 
or 

z  =  3r2-  16 


Figure  5.109  The  capsule  of 
Example  3. 


Figure  5.110  The  shadow  of 
the  capsule  in  Figure  5.109 
in  the  z  =  0  plane. 


In  addition,  the  temperature  function  may  be  described  in  cylindrical  coordi- 
nates as 


T(x,y,z)  =  z(x2  +  y2)  =  zr2. 


Hence,  we  may  calculate 


[^]avg  — 


/o*  io /2  ii-16  zrl  -rdzdr  dO 


ft ft2  fl/-ibrdzdr  de 


For  the  denominator  integral, 

-lit    r5/2  r9-r2 


/      /      /        rdzdr  d6  =         /      r  ((9  -  r2)  -  (3r2  -  16))  dr  dd 

Jo      J(i       Jlr2-16  Jo  Jo 

=  /      /      (25r  -  4r3)  dr  de 
Jo  Jo 

-f(i 

2n  /625  625 


r2-r4 


5/2 


II 


-f 

Jo 


625  625tt 

,  de  =  2jz  =   . 

8        16  7  16  8 


5.6  |  Applications  of  Integration 


This  result  agrees  with  the  volume  calculation  in  Example  6  of  §5.4,  as  it  should. 
For  the  numerator  integral,  we  compute 


p2n  p5/2  p9-r2 
JO      Jo  Jir2-U 


-2ji    r5/2    /  .2 


zr  dzdrd.9 


z=9-r2 


dr  dO 


z=3r2-16 


Thus, 


Jo  Jo 

=  /      /  ^{(9-r2)2-(3r2-\6f)drde 
Jo    Jo  z 

f2n    [5/2  r3 

=  /      /      —  (-8r4  +  78r2  -  175)  dr  dO 
Jo    Jo  2 

J    f2x  p5/2 

=  o/      /     (-^7  +  78r5  -  I75r3)dr  d9 
2  Jo  Jo 

=  2/ 


_r8  +  13r6  _  121r4 

4 


2?r    15625  15625 

 d6  =  TV. 

256  256 


[^]avg  — 


■156257T/256  25 


625^/8 


32 


Center  of  Mass:  The  Discrete  Case   

Consider  a  uniform  seesaw  with  two  masses  mj  and  m2  placed  on  either  end.  If 
we  introduce  a  coordinate  system  so  that  the  fulcrum  of  the  seesaw  is  placed  at 
the  origin,  then  the  situation  looks  something  like  that  shown  in  Figure  5.111. 
Note  that  x%  <  0  <  jci.  The  seesaw  balances  if 

m\X\  +  mjX2  =  0. 

In  this  case,  the  center  of  mass  (or  "balance  point")  of  the  system  is  at  the  origin. 

But  now  suppose  m\X\  +  m2X2  ^  0.  Then  where  is  the  balance  point?  Let  us 
denote  the  coordinate  of  the  balance  point  by  x.  Before  we  find  it,  we'll  introduce 
a  little  terminology.  The  product  m, Xj  (in  this  case,  for  i  =  1, 2)  of  mass  and 
position  is  called  the  moment  of  the  ith  body  with  respect  to  the  origin  of  the 
coordinate  system.  The  sum  m  1X1  +  1112x2  is  called  the  total  moment  with  respect 
to  the  origin.  To  find  the  center  of  mass,  we  use  the  following  physical  principle, 
which  tells  us  that  a  system  of  several  point  masses  is  physically  equivalent  (in 
terms  of  moments)  to  a  system  with  a  single  point  mass. 


Guiding  physical  principle.  The  center  of  mass  is  the  point  such  that,  if  all 
the  mass  of  the  system  were  concentrated  there,  the  total  moment  of  the  new 
system  would  be  the  same  as  that  for  the  original  system. 


Putting  this  principle  into  practice  in  our  situation,  we  see  that  total  mass  M 
of  our  system  is  mi  +  m,2.  If  x  is  the  center  of  mass,  then  the  guiding  principle 
tells  that 


Mx  =  m\X\  +  M2X2- 


Chapter  5  |  Multiple  Integration 


m1  m2  m3 


X-i 


0 


Figure  5.112  A  system  of  « 
masses  distributed  on  a  line. 


ml<t 


(*i>Vi) 


(*3>y3) 


Figure  5.113  A  system  of  « 
masses  in  R2. 


That  is,  the  total  moment  of  the  new  (concentrated)  system  is  the  same  as  the  total 
moment  of  the  original  system.  Hence, 

m\X\  +  1H2X2 


X  = 


If  we  have  a  system  of  n  masses  distributed  along  a  (coordinatized)  line,  then 
the  same  reasoning  may  be  applied.  (See  Figure  5.1 12.)  We  have 


total  moment     ni\X\  +  1712X2  +  •  •  •  +  mnxn      Yl"=i  mixi 


total  mass 


mi  +  m2  H  h  m„ 


(2) 


Now  we  move  to  two  and  three  dimensions.  Suppose,  first,  that  we  have  n 
particles  (or  bodies)  arranged  in  the  plane  as  in  Figure  5.1 13.  Then  there  are  two 
moments  to  consider: 

n 

Total  moment  with  respect  to  the  y-axis  =       m,*, , 


1=1 


and 


Total  moment  with  respect  to  the  x-axis  =  m, -y,- . 


1=1 

(Admittedly,  this  terminology  may  seem  confusing  at  first.  The  idea  is  that  the 
moment  measures  how  the  system  balances  with  respect  to  the  coordinate  axes. 
It  is  the  x-coordinate — not  the  y-coordinate — that  measures  position  relative  to 
the  y-axis.  Similarly,  the  y-coordinate  measures  position  relative  to  the  x-axis.) 
The  guiding  principle  tells  us  that  the  center  of  mass  is  the  point  (x ,  y)  such  that, 
if  all  the  mass  of  the  system  were  concentrated  there,  then  the  new  system  would 
have  the  same  total  moments  as  the  original  system.  That  is,  if  M  =  ^  m;,  then 


Mx  =  rriiXi 


(i.e.,  the  moment  with  respect  to  the  y-axis  of  the  new  system  equals  the  moment 
with  respect  to  the  y-axis  of  the  original  system)  and 


My  =  y^m,y,-. 
Thus,  we  have  shown  the  following: 


Discrete  center  of  mass  in  R2.  Given  a  system  of  n  point  masses  m\, 
m2, . . . ,  m„  at  positions 

(xi,yi),    (x2,  y2)   (x„,yn)  inR2, 

the  coordinates  (x ,  y)  of  the  center  of  mass  are 


5.6  |  Applications  of  Integration  379 

For  particles  arranged  in  three  dimensions,  little  more  is  needed  than  adding 
an  additional  coordinate.  (See  Figure  5.1 14.) 


Discrete  center  of  mass  in  R  .  Given  a  system  of  n  point  masses  m\, 
ni2,  ■ .  ■ ,  mn  at  positions 

(xuyi,zi),    (x2,  y2,  Zi),  ■  ■  - ,    (x„,y„,z„)  inR3, 

the  coordinates  (x,  y,  z)  of  the  center  of  mass  are  given  by 

*=  ,    y=^'    — ,    and    z=^'      '■  (4) 


The  numerators  of  the  fractions  in  (4)  are  the  moments  with  respect  to  the 
coordinate  planes.  Thus,  for  example,  the  sum  Yl"=i  mix>  is  me  total  moment 
with  respect  to  the  yz-plane. 

By  definition,  moments  of  physical  systems  are  additive.  That  is,  the  total 
moment  of  a  system  is  the  sum  of  the  moments  of  its  constituent  pieces.  However, 
it  is  by  no  means  the  case  that  a  coordinate  of  the  center  of  mass  of  a  system  is 
the  sum  of  the  coordinates  of  the  centers  of  mass  of  its  pieces.  This  additivity 
property  makes  the  study  of  moments  important  in  its  own  right. 

Center  of  Mass:  The  Continuous  Case  

Now,  we  turn  our  attention  to  physical  systems  where  mass  is  distributed  in 
a  continuous  fashion  throughout  the  system  rather  than  at  only  finitely  many 
isolated  points. 

To  begin  with  the  one-dimensional  case,  suppose  we  have  a  straight  wire 
placed  on  a  coordinate  axis  between  points  x  =  a  and  x  =  b  as  shown  in 
Figure  5.115.  Moreover,  suppose  that  the  mass  of  this  wire  is  distributed  according 
to  some  continuous  density  function  <5  (x ) .  We  seek  the  coordinate  x  that  represents 
the  center  of  mass,  or  "balance  point,"  of  the  wire. 


xn=b 

Figure  5.1 1 5  A  "coordinatized"  wire. 
The  mass  of  the  segment  between  jc;_i  and 
Xi  is  approximately  8(x*)Axj. 

Imagine  breaking  the  wire  into  n  small  pieces.  Since  the  density  is  continuous, 
it  will  be  nearly  constant  on  each  small  piece.  Thus,  for  i  =  1 , . . . ,  n,  the  mass  m, 
of  each  piece  is  approximately  8(x*)Ax, ■  ,  where  A  the  length  of 

each  segment  of  wire,  and  x*  is  any  number  in  the  subinterval  [jc,_i  ,  #,•].  Hence, 
the  total  mass  is 

n  n 

M  =  J2'ni  ~  ^2s(x*)Axi, 

i=l  i=l 

and  the  total  moment  with  respect  to  the  origin  is  approximately 

xf  8(x*)Ax,. 

approx.  approx. 
position  mass 


mi 


-  y 


X 

Figure  5.1 1 4  A  discrete  system 
of  masses  in  R3. 


J- — I  1  1 — H- 

d  —  Xq   X\      X-j      ■  ■  ■       X;  _  i  X; 


380       Chapter  5  |  Multiple  Integration 


Of  course,  these  results  can  be  used  to  provide  an  approximation  of  the  coordi- 
nate x  of  the  center  of  mass.  For  an  exact  result,  however,  we  let  all  the  pieces 
of  wire  become  "infmitesimally  small";  that  is,  we  take  limits  of  the  foregoing 
approximating  sums  as  all  the  Ax,  's  tend  to  zero.  Such  limits  give  us  integrals, 
and  we  may  reasonably  define  our  terms  as  follows: 


Continuous  center  of  mass  in  R.  For  a  wire  located  along  the  x-axis 
between  x  =  a  and  x  =  b  with  continuous  density  per  unit  length  <5(x): 


f 


Total  mass  =  /  8(x)dx 


Total  moment  =  /   xS(x)dx.  (5) 

J  a 

total  moment      f  xS(x)dx 

Center  of  mass  x  = 


total  mass        f*  8(x)dx 


Compare  the  formulas  in  (3)  with  those  in  (5).  Instead  of  a  sum  of  masses 
and  a  sum  of  products  of  mass  and  position,  we  have  an  integral  of  "infinitesimal 
mass"  (the  <5(x)  dx  term)  and  an  integral  of  infinitesimal  mass  times  position. 


EXAMPLE  4  Suppose  that  a  wire  is  located  between  x  =  —  1  and  x  =  1  along 
a  coordinate  line  and  has  density  S(x)  =  x2  +  1.  Using  the  formulas  in  (5),  we 
compute  that  the  center  of  mass  has  coordinate 


Figure  5.116  A  lamina 
depicted  as  a  region  D  in 
the  xy-plane  with  density 
function  S. 


f\x{x2+\)dx  _  {jx^+jx^ti 
f\(x2+l)dx  (ix3+x)|lj 


4  =  0. 

3 


This  makes  sense,  since  this  wire  has  a  symmetric  density  pattern  with  respect  to 
the  origin  (i.e.,  <5(x)  =  S(—x)).  ♦ 

The  analogous  situation  in  two  dimensions  is  that  of  a  lamina  or  flat  plate  of 
finite  extent  and  continuously  varying  density  S(x,  y).  (See  Figure  5.116.)  Using 
reasoning  similar  to  that  used  to  obtain  the  formulas  in  (5),  we  make  the  following 
definition  for  the  coordinates  (x,  y)  of  the  center  of  mass  of  the  lamina: 


Continuous  center  of  mass  in  R2.  For  a  lamina  represented  by  the  region 
D  in  the  xy-plane  with  continuous  density  per  unit  area  8(x,  y): 


x  = 


total  moment  with  respect  to  y-axis      ffDx  S(x,  y)dA 

'   JfDS(x,y)dA  ; 

ffDyS(x,y)dA 


total  mass 

total  moment  with  respect  to  x-axis 
total  mass 


(6) 


ffDS(x,y)dA  ■ 


5.6  |  Applications  of  Integration 


Roughly,  the  term  S(x,  y)dA  represents  the  mass  of  an  "infinitesimal  two-dimen- 
sional" piece  of  the  lamina  and  the  various  double  integrals  the  limiting  sums  of 
such  masses  or  their  corresponding  moments. 

EXAMPLE  5  We  wish  to  find  the  center  of  mass  of  a  lamina  represented  by  the 
region  D  in  R2  whose  boundary  consists  of  portions  of  the  parabola  y  =  x2  and 
the  line  y  =  9  and  whose  density  varies  as  S(x,  y)  =  x2  +  y.  (See  Figure  5.117.) 

First,  note  that  this  lamina  is  symmetric  with  respect  to  the  y-axis  and  that,  in 
addition,  the  density  function  has  a  similar  symmetry  because  8 (x,  y)  =  S(—x,  y). 
We  may  conclude  from  these  two  observations  that  the  center  of  mass  must  occur 
along  the  y-axis  (i.e.,  that  x.  =  0).  Using  the  formulas  in  (6)  and  noting  that  the 
lamina  is  represented  by  an  elementary  region  of  type  1 , 


y  = 


JfDyS(x,y)dA  _  j%  f^yjx2  +  y)dy  dx 
ffDS(x,y)dA   :  :  /\ •  y)dydx 


For  the  denominator  integral,  we  compute 


x  +  y)  dy  dx  = 


dx 


For  the  numerator, 


9 


y(x2  +  y)dy  dx  = 


dx 


x6  x6\l 

 1  dx 

2       3  )\ 


3 


11664 


7 


Hence, 


11664/7  _  45 
1296/5  ~  T 


6.43. 


This  answer  is  quite  plausible,  since  the  density  of  the  lamina  increases  with  y, 
and  so  we  should  expect  the  center  of  mass  to  be  closer  to  y  =  9  than  to  y  =  0. 

♦ 


382       Chapter  5  |  Multiple  Integration 


We  may  modify  the  two-dimensional  formulas  to  produce  three-dimensional 
ones. 


Continuous  center  of  mass  in  R3.  Given  a  solid  W  whose  density  per 
unit  volume  varies  continuously  as  8(x,  y,  z),  we  compute  the  coordinates 
(x,  y,  z)  of  the  center  of  mass  of  W  using  the  following  quotients  of  triple 
integrals: 

total  moment  with  respect  to  yz-plane  fffwx8(x,y,z)dV 
total  mass 


fffwS(x,y,z)dV  ' 


y 


total  moment  with  respect  to  xz-plane     fffwy  Kxi  y,  z)dV 


total  mass  fffw  8(x,  y,  z)dV 

total  moment  with  respect  to  x  y-plane      fffwz  S(x,  y,z)dV 
total  mass  fffwS(x,y,z)dV 


-;  (7) 


(0,0,3) 


Plane 

x  +  y  +  z =3 


(3,0, 0) 


Figure  5.118 

Example  6. 


The  tetrahedron  of 


In  (7)  we  may  think  of  the  term  S(x,  y,  z)dV  as  representing  the  mass  of  an 
"infinitesimal  three-dimensional"  piece  of  W.  Then  the  triple  integrals  are  the 
limiting  sums  of  masses  or  moments  of  such  pieces. 

EXAMPLE  6  Consider  the  solid  tetrahedron  W  with  vertices  at  (0,  0,  0), 
(3,  0,  0),  (0,  3,  0),  and  (0,  0,  3).  Suppose  the  mass  density  at  the  point  (x,  y,  z) 
inside  the  tetrahedron  is  8(x,  y,z)  =  x  +  y  +  z+  1.  We  calculate  the  resulting 
center  of  mass.  (See  Figure  5.118.) 

First,  note  that  the  position  of  the  tetrahedron  in  space  and  the  density  function 
are  both  such  that  the  roles  of  x,  y,  and  z  may  be  interchanged  freely.  Hence,  the 
coordinates  (x,  y,  z)  of  the  center  of  mass  must  satisfy  x  —  y  —  z.  Therefore,  we 
may  reduce  the  number  of  calculations  required. 

The  tetrahedron  is  a  type  4  elementary  region  in  space.  Thus,  we  may  calculate 
the  total  mass  M  of  W,  using  the  following  iterated  integral: 


l-x  r3-x-y 


M 


if  I 

Jo  Jo  Jo 
fo  fo  \ 
Jo  Jo 

=  f  m- 

Jo 


(x  +  y  +  z  +  l)dzdy  dx 

Z=3—x—y 


(x  +  y  +  \)z  + 


1T2 

2X 


y  -xy 


dy  dx 


2  y2)  dy  dx 


x2)y-\{\+x)y2-\y% 


1-2 

2 


=3-.i 


dx 


The  total  moment  with  respect  to  the  xy-plane  is  given  by 

O    r3-x  i-3-x-y 


ITI 

Jo  Jo  Jo 


z(x  +  y  +  z  +  1)  dz  dy  dx 


5.6  |  Applications  of  Integration 


•3    ,3-,   /  2  z3 

(x  +  y+l)-  +  j 


If 

Jo  Jo 

/T"< 

Jo  Jo 


=3-x-y 


dy  dx 


2         2X  +  + 


+  xy+  \x2y  +  \y2  +  \xy2  +  iy3)  dy  dx 


Jo 


117  27, 


15  „2  _  1  v3 
4  A  6A 


lx4)dx=459 


Hence, 


x  =  y  =  z  = 


459 
40 


117 


51 

65 


24 


0.7846. 


40  ■ 


If  an  object  is  uniform,  in  the  sense  that  it  has  constant  density,  then  one  uses 
the  term  centroid  to  refer  to  the  center  of  mass  of  that  object.  Suppose  the  object 
is  a  solid  region  W  in  R3.  Then,  if  the  density  <5  is  a  constant  k,  the  equations  for 
the  coordinates  (x,y,z)  may  be  deduced  from  those  in  (7).  For  the  x-coordinate, 
we  have 


x  = 


fffwxS(x,y,z)dV  =  fffwkxdV 
fffw8(x,y,z)dV  fffwkdV 

fffwxdV  _  1 
fffwdV      volume  of  W 


1 1  Iw 


xdV. 


Similarly, 
y  = 


1 


volume  of  W 


f  f  fy 


ydV    and  z=  

volume  of  W 


ffl 


zdV. 


In  particular,  the  constant  density  S  plays  no  role  in  the  calculation  of  the 
centroid,  only  the  geometry  of  W.  (Note:  Completely  analogous  statements  can 
be  made  in  the  case  of  centroids  of  laminas  in  R2.) 


Figure  5.1 19  The  cone  of 
Example  7. 


EXAMPLE  7  We  compute  the  centroid  of  a  cone  of  radius  a  and  height  h .  (See 
Figure  5.119.) 

By  symmetry,  x  =  y  =  0.  Moreover,  we  know  that  the  volume  of  the  cone  is 
(ir/3)a2h.  Thus,  the  z-coordinate  of  the  centroid  is 


=  —111 

na2h  J  J  Jw 


zdV. 


This  triple  integral  is  most  readily  evaluated  by  using  cylindrical  coordinates. 
(See  Example  15  of  §5.5.)  The  lateral  surface  of  the  cone  is  given  by  z  =  \r,  so 
we  calculate 


^        /,2jr  pa 

=  — 7T-  I      I     I    zrdzdrd0  = 
nazh  J0    Jo  J'ir 


Tca2h 


jta2h2\  3 

  I  =  -h 

4/4 


after  a  straightforward  evaluation.  Hence,  the  centroid  of  the  cone  is  located  at 
(0,0,  \h).  ♦ 


384       Chapter  5  |  Multiple  Integration 


Moments  of  Inertia   

Let  W  be  a  rigid  solid  body  in  space.  As  we  have  seen,  the  moment  integral 
with  respect  to  the  xy-plane  is  Mxy  =  fffwz  8(x,  y,  z)dV — that  is,  the  inte- 
gral of  the  product  of  the  position  relative  to  a  reference  plane  (in  this  case  the 
xy-plane)  and  the  density  of  the  solid.  This  integral  can  be  considered  to  mea- 
sure the  ease  with  which  W  can  be  displaced  perpendicularly  from  the  reference 
plane. 

Now,  consider  spinning  W  about  a  fixed  axis  (which  may  or  may  not  pass 
through  W).  The  moment  of  inertia  /  (or  second  moment — the  moment  integral 
mentioned  in  the  preceding  paragraph  is  sometimes  called  the  first  moment)  is 
a  measure  of  the  ease  with  which  W  can  be  made  to  spin  about  the  given  axis. 
Specifically,  /  is  the  integral  of  the  product  of  the  density  at  a  point  in  W  and  the 
square  of  the  distance  from  that  point  to  a  fixed  axis;  that  is, 


1 1  Iw 


dl8{x,y,z)dV,  (8) 

w 

where  d  is  the  distance  from  (x,  y,  z)  6  W  to  the  specified  axis. 
When  the  axes  of  rotation  are  the  coordinate  axes  in  R3,  we  have 


/v  =  moment  of  inertia  about  the  x-axis  =  /  /  /  (y  +  z  )8(x,  y,  z)  dV; 

'w 


ffl 

Iy  =  moment  of  inertia  about  the  y-axis  =  /  /  /  (x2  +  z2)  S(x,  y,  z)  d  V; 

J  J  Jw 

I-  =  moment  of  inertia  about  the  z-axis  =  /  /  /  (x2  +  y2)S(x,  y,  z)dV. 

J  J  Jw 


Figure  5.120  The  box  of 

Example  8. 


EXAMPLE  8   Let  W  be  a  solid  box  of  uniform  density  S  and  dimensions  a,  b, 
and  c.  If  W  is  situated  symmetrically  with  respect  to  the  coordinate  axes  as  shown 
in  Figure  5.120,  we  compute  the  moments  of  inertia  with  respect  to  these  axes. 
Note,  first,  that  W  may  be  described  as 


W  =  \(x,y,z) 


a  a      b  be  c 

-  <  x  <  -,  —  <  y  <  —  <  z  <  - 
2  ~     ~  2      2~  y  ~  2      2  ~  "2 


Hence,  the  moment  of  inertia  about  the  x-axis  is 

/c/2     rb/2    pa /2  pc/2  pb/2 

/       /     (v2  +  z2)Sdxdydz=  /       /     (y2  +  z2)Sadydz 
-c/2  J-b/2  J -a/2  J -c/2  J-b/2 

fc/2    /   3  \   y=bl2  fc /2   /b3  \ 

=SaLS^+zy),---Jz=SaLM+bz) 


+  bzz  dz 


,'b^c  bc^ 

Sal  1  

12  12 


Sabc 


12 


'-(b2  +  c2). 


5.6  |  Applications  of  Integration 


By  permuting  the  roles  of  x,  y,  and  z  (and  the  corresponding  constants  a,  b,  and 
c),  we  see  that 

Sabc    0      i  Sabc    ,  , 

ly  =  —{a2  +  c2)    and    L  =  —{a2  +  b2). 

Therefore,  if  a  >  b  >  c  (as  in  Figure  5.120),  it  follows  that  Ix  <  Iy  <  /,.  This 
result  may  be  confirmed  by  the  observation  that  rotations  about  the  axis  parallel 
to  the  longest  side  of  the  box  are  easiest  to  effect  in  that  the  same  torque  applied 
about  each  axis  will  cause  the  most  rapid  rotation  to  occur  about  the  axis  through 
the  longest  dimension.  A  related  fact  is  regularly  exploited  by  figure  skaters  who 
pull  their  arms  in  close  to  their  bodies,  thereby  reducing  their  moments  of  inertia 
and  speeding  up  their  spins.  ♦ 

EXAMPLE  9  Let  W  be  the  solid  bounded  by  the  cone  z  =  2^/x2  +  y2  and  the 
plane  z  =  4  shown  in  Figure  5.121.  Assume  that  the  density  of  material  inside  W 
varies  as  8(x,  y,  z)  =  5  —  z.  Let  us  calculate  the  moment  of  inertia  Iz  about  the 
z-axis. 

Given  the  geometry  of  the  situation,  it  is  easiest  to  work  in  cylindrical  coor- 
dinates, in  which  case  the  cone  is  given  by  the  cylindrical  equation  z  =  2r.  Thus, 
we  have 

L  =  f  [  [  (x2  +  y2)S(x,y,z)dV  =  f     f    f  r2(5  -  z)rdzdrd6 

J  J  Jw  Jo       Jo  Jlr 

=         /    r3  (5z  ~  \z2)\Z~2r  drM  =  /      /    (12r3  "  10r4  +  2r5)  drde 
Jo    Jo  Jo  Jo 

f2n  ,    i        s     1  /r\  i r=2  f2lT  16  32;r 

Jo  Jo 


Recall  that  the  center  of  mass  of  a  solid  object  of  total  mass  M  is  the  point 
such  that  if  all  the  mass  M  were  concentrated  there,  the  (first)  moment  would 
remain  the  same.  An  analogous  idea  may  be  defined  in  the  context  of  moments  of 
inertia.  The  radius  of  gyration  of  a  solid  with  respect  to  an  axis  is  the  distance 
r  from  that  axis  that  we  should  locate  a  point  of  mass  M  so  that  it  has  the  same 
moment  of  inertia  /  (with  respect  to  the  axis)  as  the  original  solid  does.  More 
concisely,  the  radius  of  gyration  r  is  defined  by  the  equation 


EXAMPLE  10  We  determine  the  radius  of  gyration  with  respect  to  the  z-axis 
of  the  cone  described  in  Example  9.  Hence,  we  compute 


386       Chapter  5  |  Multiple  Integration 


From  Example  9,  L  =  32n/3.  We  determine  the  total  mass  M  of  the  cone  as 
follows: 

M=         /    /  (5  -  z)rdzdrd6  =         /  (12  -  lOr  +  2r2)r  dr  d9 

J0      J0    Jlr  Jo  Jo 


=r((*i-^+Mi^*=r^=f- 

Thus, 

'     Y  32^/3 


5.6  Exercises 


1 .  The  local  grocery  store  receives  a  shipment  of  75  cases 
of  cat  food  every  month.  The  inventory  of  cat  food  (i.e., 
the  number  of  cases  of  cat  food  on  hand  as  a  function 
of  days)  is  given  by  I(x)  =  75  cos(jrx/15)  +  80. 

(a)  What  is  the  average  daily  inventory  over  a  month? 

(b)  If  the  cost  of  storing  a  case  is  2  cents  per  day, 
determine  the  average  daily  holding  cost  over  the 
month. 

2.  Find  the  average  value  of  f(x,  y)  =  sin2  x  cos2  y  over 
R  =  [0,2jt]  x  [0,4tt]. 

3.  Find  the  average  value  of  f(x,  y)  =  e2x+y  over  the  tri- 
angular region  whose  vertices  are  (0,  0),  (1,  0),  and 
(0,  1). 

4.  Find  the  average  value  of  g(x,  y,  z)  =  ez  over  the  unit 
ball  given  by 

B  =  {(x,y,z)  |  *2  +  y2  +  z2  <  1}. 

5.  Suppose  that  the  temperature  at  a  point  in  the  cube 

W  =  [-1,  1]  x  [-1,  1]  x  [-1,  1] 

varies  in  proportion  to  the  square  of  the  point's  distance 
from  the  origin. 

(a)  What  is  the  average  temperature  of  the  cube? 

(b)  Describe  the  set  of  points  in  the  cube  where  the 
temperature  is  equal  to  the  average  temperature. 

6.  Let  D  be  the  region  between  the  square  with  ver- 
tices (1,  1),  (-1,  1),  (-1,  -1),  (1,  -1)  and  the  unit 
disk  centered  at  the  origin.  Find  the  average  value  of 
f(x,  y)  =  x2  +  y2  on  D. 

7.  Let  W  be  the  region  in  R3  between  the  cube  with 
vertices  (1,  1,  1),  (-1,  1,  1),  (-1,  -1,  1),  (1,  -1,  1), 
(1,1,-1),  (-1,1,-1),  (-1,-1,-1),  (1,-1,-1) 
and  the  unit  ball  centered  at  the  origin.  Find  the  av- 
erage value  of  f(x,  y,  z)  =  x2  +  y2  +  z2  on  W. 


8.  Suppose  that  you  commute  every  day  to  work  by  sub- 
way. You  walk  to  the  same  subway  station,  which  is 
served  by  two  subway  lines,  both  stopping  near  where 
you  work.  During  rush  hour,  each  subway  line  sends 
trains  to  arrive  at  the  stop  every  6  minutes,  but  the  dis- 
patchers begin  the  schedules  at  random  times.  What  is 
the  average  time  you  expect  to  wait  for  a  subway  train? 
(Hint:  Model  the  waiting  time  for  the  two  subway  lines 
by  using  a  point  (x,  y)  in  the  square  [0,  6]  x  [0,  6].) 

9.  Repeat  Exercise  8  in  the  case  that  the  subway  stop  is 
serviced  by  three  subway  lines  (each  with  trains  arriv- 
ing every  6  minutes),  rather  than  two. 

10.  Find  the  center  of  mass  of  the  region  bounded  by  the 
parabola  y  =  8  —  2x2  and  the  *-axis 

(a)  if  the  density  S  is  constant; 

(b)  if  the  density  S  =  3y. 

11.  Find  the  centroid  of  a  semicircular  plate.  (Hint:  Judi- 
cious use  of  a  suitable  coordinate  system  might  help.) 

1 2.  Find  the  center  of  mass  of  a  plate  that  is  shaped  like  the 
region  between  y  =  x2  and  y  =  2x,  where  the  density 
varies  as  1  +  x  +  y. 

13.  Find  the  center  of  mass  of  a  lamina  shaped  like  the 
region 

{(*,  y)  |  0  <  y  <  *Jx,  0  <  x  <  9}, 
where  the  density  varies  as  xy. 

1 4.  Find  the  centroid  of  the  region  bounded  by  the  cardioid 
given  in  polar  coordinates  by  the  equation  r  = 
1  —  sin  9.  (Hint:  Think  carefully.) 

15.  Find  the  centroid  of  the  lamina  described  in  polar 
coordinates  as 

{(r,9)  |  0  <  r  <  4cos0,  0  <  6  <  tt/3}. 


5.6  I  Exercises  387 


1 6.  Find  the  center  of  mass  of  the  lamina  described  in  polar 
coordinates  as 

l(r,6)  |  0  <  r  <  3,  0  <6  <  tt/4}, 

where  the  density  of  the  lamina  varies  as  S(r,  0)  = 
4-r. 

17.  Find  the  center  of  mass  of  the  region  inside  the  car- 
dioid  given  in  polar  coordinates  as  r  =  1  +  cos  8,  and 
whose  density  varies  as  S(r,  6)  =  r. 

1 8.  Find  the  centroid  of  the  tetrahedron  whose  vertices  are 
at  (0,  0,  0),  (1,  0,  0),  (0,  2,  0),  and  (0,  0,  3). 

19.  A  solid  is  bounded  below  by  z  =  3y2,  above  by  the 
plane  2  =  3,  and  on  the  ends  by  the  planes  x  =  —  1 
and  x  =  2. 

(a)  Find  the  centroid  of  this  solid. 

(b)  Now  assume  that  the  density  of  the  solid  is  given 
by  S  =  z  +  x2.  Find  the  center  of  mass  of  the 
solid. 

20.  Determine  the  centroid  of  the  region  bounded  above 
by  the  sphere  x2  +  y2  +  z2  =  1 8  and  below  by  the 
paraboloid  3z  =  x2  +  y2 . 

21 .  Find  the  centroid  of  the  solid,  capsule-shaped  region 
bounded  by  the  paraboloids  z  =  3x2  +  3y2  —  16  and 

z  =  9-x2-y2. 

22.  Find  the  centroid  of  the  "ice  cream  cone"  shown  in 
Figure  5.122. 


z 

Sphere:  x2  +  y2  +  z2  =  25 


x 


Figure  5.122  The  ice  cream  cone  solid 
of  Exercise  22. 

23.  Find  the  centroid  of  the  solid  shaped  as  one-eighth  of 
a  solid  ball  of  radius  a.  (Hint:  Model  the  solid  as  the 
first  octant  portion  of  a  ball  of  radius  a  with  center  at 
the  origin. ) 

24.  Find  the  center  of  mass  of  a  solid  cylindrical  peg  of 
radius  a  and  height  h  whose  mass  density  at  a  point 
in  the  peg  varies  as  the  square  of  the  distance  of  that 
point  from  the  top  of  the  cylinder. 


25.  (a)  Find  the  moment  of  inertia  about  the  coordinate 

axes  of  a  solid,  homogeneous  tetrahedron  whose 
vertices  are  located  at  (0,0,0),  (1,  0,  0),  (0,  1,  0), 
and(0,  0,  1). 

(b)  What  are  the  radii  of  gyration  about  the  coordinate 

axes? 

26.  Consider  the  solid  cube  W  =  [0,  2]  x  [0,  2]  x  [0,  2]. 
Find  the  moments  of  inertia  and  the  radii  of  gyration 
about  the  coordinate  axes  if  the  density  of  the  cube  is 
S(x,  y,  z)  =  x  +  y  +  z  +  1. 

27.  A  solid  is  bounded  by  the  paraboloid  z  =  x2  +  y2  and 
the  plane  z  =  9.  Find  the  moment  of  inertia  and  radius 
of  gyration  about  the  z-axis  if 

(a)  the  density  is  S(x,  y,  z)  =  2z; 

(b)  the  density  is  S(x,  y,  z)  =  *Jx2  +  y2. 

28.  Find  the  moment  of  inertia  and  radius  of  gyration  about 
the  z-axis  of  a  solid  ball  of  radius  a,  centered  at  the  ori- 
gin, if 

(a)  the  density  S  is  constant; 

(b)  S(x,y,z)  =  x2  +  y2  +  z2; 

(c)  S(x,y,z)  =  x2  +  y2. 

We  can  find  the  moment  of  inertia  of  a  lamina  in  the  plane 
with  density  S(x,  y)by  considering  the  lamina  to  be  a  flat  plate 
sitting  in  the  xy -plane  in  R3.  Then,  for  example,  the  distance 
of  a  point  (x,  y)  in  the  lamina  to  the  x-axis  is  given  by  \y\,  the 
distance  to  the  y-axis  is  given  by  \x\,  and  the  distance  to  the 

z-axis  (or  the  origin)  is  given  by  *J x2  +  y2.  (See  Figure  5.123.) 
Using  these  ideas,  find  the  specified  moments  of  inertia  of  the 
laminas  given  in  Exercises  29—31. 


z 


29.  The  moment  of  inertia  Ix  about  the  x-axis  of  the 
lamina  that  has  the  shape  bounded  by  the  graph  of 
y  =  x2  +2  and  the  line  y  =  3,  and  whose  density 
varies  as  S(x,  y)  =  x2  +  1. 

30.  The  moment  of  inertia  /,  about  the  z-axis  of  the  lam- 
ina shaped  as  the  rectangle  [0,  2]  x  [0,  1],  and  whose 
density  varies  as  S(x,  y)  =  1  +  y. 


388       Chapter  5  |  Multiple  Integration 


31 .  The  moment  of  inertia  about  the  line  y  =  3  of  the  lam- 
ina shaped  as  the  disk 

{(x,y)\x2  +  y2  <4}, 

and  whose  density  varies  as  S(x,  y)  =  x2. 

The  gravitational  field  between  a  mass  M  concentrated  at 
the  point  (x,  y,  z)  and  a  mass  m  concentrated  at  the  point 
(xo,  Jo,  Zo)  is 

p_     GMm[(x  -  x0)i  +  (y  -  ypjj  +  (z  -  z0)k] 
[(x  -  x0)2  +  (y  -  y0)2  +  (z  -  za)2fl2 
The  gravitational  potential  V  ofF  is 

_  GMm 

y/(x  -x0)2  +  (y-  y0)2  +  (z  -  Z0)2  ' 

(We  have  seen  in  §3.3  that  F  =  -VV.J  Now  suppose  that,  in- 
stead of  a  point  mass  M,  we  have  a  solid  region  W  of  density 
S(x,  y,  z)  and  total  mass  M.  Then  the  gravitational  potential  of 
W  acting  on  the  point  mass  m  may  be  found  by  looking  at  "in- 
finitesimal "point  masses  dm  =  S(x,  y,z)dV  and  adding  (via 
integration)  their  individual  potentials.  That  is,  the  potential 
ofW  is 

V(x0,  y0,  z0) 

err  GmS(x,y,z)dv 

J  J  Jw  J{x  -  x0)2  +  (y  -  y0)2  +  (z  -  z0)2 ' 

In  Exercises  32-34,  Let  W  be  the  region  between  two  con- 
centric spheres  of  radii  a  <  b,  centered  at  the  origin.  (See 
Figure  5.124.)  Assume  that  W  has  total  mass  M  and  constant 
density  8.  The  object  of  the  following  exercises  is  to  compute 
the  gravitational  potential  V(xo,  yo,  Zo)  of  W  on  a  mass  m 
concentrated  at  (xo,  yo,  Zo)-  Note  that,  by  the  spherical  sym- 
metry, there  is  no  loss  of  generality  in  taking  (xq,  yo,  Zo)  equal 


to  (0,  0,  r).  So,  in  particular,  r  is  the  distance  from  the  point 
mass  m  to  the  center  ofW. 


Figure  5.124  The  spherical 
shell  of  Exercises  32-34. 


32.  Show  that  if  r  >  b,  then  V(0,  0,  r)  =  -GMm/r.  This 
is  exactly  the  same  gravitational  potential  as  if  all  the 
mass  M  of  W  were  concentrated  at  the  origin.  This  is  a 
key  result  of  Newtonian  mechanics.  (Hint:  Use  spher- 
ical coordinates  and  integrate  with  respect  to  <p  before 
integrating  with  respect  to  p.) 

33.  Show  that  if  r  <  a,  then  there  is  no  gravitational  force. 
(Hint:  Show  that  V(0,  0,  r)  is  actually  independent  of  r. 
Then  relate  the  gravitational  potential  to  gravitational 
force.  As  in  Exercise  32,  use  spherical  coordinates  and 
integrate  with  respect  to  (p  before  integrating  with  re- 
spect to  p.) 

34.  (a)  Find  V(0,  0,  r)  if  a  <  r  <  b. 

(b)  Relate  your  answer  in  part  (a)  to  the  results  of  Ex- 
ercises 32  and  33. 


5.7   Numerical  Approximations  of  Multiple 
Integrals  (optional) 

Suppose  that  /  is  a  continuous  function  defined  on  some  bounded  region  D  in  the 
plane.  If  D  is  a  rectangle  or  an  elementary  region,  then  Theorem  2. 1 0  enables  us  to 
calculate  ffDf  dA  as  an  iterated  integral.  However,  the  "partial  antiderivatives" 
in  the  iterated  integral  may  turn  out  to  be  impossible  to  determine  in  practice. 
Nonetheless,  an  approximate  value  of  ffD  f  dA  may  be  entirely  acceptable  for 
some  purposes.  In  this  section,  we  discuss  how  we  can  adapt  numerical  meth- 
ods for  approximating  definite  integrals  of  functions  of  a  single  variable  to  give 
techniques  for  approximating  double  integrals. 

Numerical  Methods  for  Definite  Integrals   

Let  us  review  two  familiar  techniques  for  approximating  the  value  of  fa  f(x)  dx. 
The  first  of  these  is  the  trapezoidal  rule.  We  begin  by  partitioning  the  interval 
[a,  b]  into  n  equal  subintervals.  Thus,  we  set 

b  —  a 

Ax  =   ,     a  =  xo  <  x\  <  . . .  <  xn  =  b,  where  Xj  =  a  +  i  Ax. 

n 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


The  trapezoidal  rule  approximation  T„  is 

Tn  =  ^  U(a)  +  2 fix,)  +■■■+  2/(x»_i)  +  f(b)] 

(1) 

It  is  obtained  by  approximating  the  function  /  with  n  linear  functions  that  pass 
through  the  pairs  of  points  (x,_i ,  /(x,_i))  and  (x,- ,  /(*,•))  for  r  =  1, . . . ,  n.  Then 
the  (net)  area  under  the  curve  y  =  f(x)  between  x  =  a  andx  =  b  is  approximated 
by  the  sum  of  the  (net)  areas  under  the  graphs  of  the  linear  functions.  When  / 
is  nonnegative,  the  approximating  areas  are  those  of  trapezoids — thus,  the  name 
for  the  method.  (See  Figure  5.125.) 

The  key  theoretical  result  concerning  the  trapezoidal  rule  is  given  by  the 
following. 


Ax 
~2~ 


n-l 


/(a) +  2  £/(*,)  +  /(&) 


THEOREM  7.1    Given  a  function  /  that  is  integrable  on  [a,  b],  we  have 


/' 

J  a 


f{x)dx  =  Tn  +  E„, 


where 


Ax 
~1~ 


n-l 


f(a)  +  2j2f(xi)  +  f(t>) 


and  E„  denotes  the  error  involved  in  using  Tn  to  approximate  the  value  of  the 
definite  integral.  In  addition,  if  /  is  of  class  C2  on  [a,  b],  then  there  is  some 
number  f  in  (a ,  b)  such  that 


F  — 


(Ax)2/"(0  = 


12 


12n2 


In  particular,  Theorem  7.1  shows  that  if  /"  is  bounded  that  is,  if  \f"(x)\  <  M 
for  all  x  in  [a,  b],  then  \En\  <  (b  —  a)3 M/(l2n2).  This  inequality  is  very  useful 
in  estimating  the  accuracy  of  the  approximation. 

EXAMPLE  1  We  approximate  f0  sin(x2)  dx  using  the  trapezoidal  rule  with 
n  =  4.  Thus,  we  have 

1  -  0 

Ax  =  =  0.25, 

4 

so  that,  using  (1),  we  have 


I 


1  0  25 

sin(x2)Jx  R«  T4  =  —  [sin0  +  2sin(0.252)  +  2sin(0.52)  +  2sin(0.752) 


+  sinl] 
0.25 


2 
0.25 


[0  +  0.12492  +  0.49481  +  1.06661  +  0.84147] 
[2.52780]  =  0.31598. 


Chapter  5  |  Multiple  Integration 


Note  that  the  second  derivative  of  sin(x2)  is  2  cos  (x2)  —  Ax2  sin(x2)  so  that,  for 
0  <  x  <  1, 

|2cos(x2)-4x2sin(x2)|  <  2|  cos(x2)|  +  4x2|  sin(x2)|  <  2  +  4  =  6. 
Hence,  using  Theorem  7.1, 

(1  -  0)36  1 

|£4|  <  -  V-  =  —  =  0.03125. 

1     1  "    12 -42  32 

Thus,  the  true  value  ofthe  integral  must  be  between  0.3 1598  -  0.03125  =  0.2847 
(rounding  to  four  decimal  places)  and  0.31598  +  0.03125  =  0.3473.  Of  course, 
a  more  accurate  approximation  may  be  obtained  by  taking  a  finer  partition  (i.e., 
a  larger  value  for  n).  ♦ 

Another  numerical  technique  for  approximating  the  value  of  f(x)dx  with 
which  you  may  be  familiar  is  Simpson's  rule.  As  with  the  trapezoidal  rule,  we 
partition  the  interval  [a,  b]  into  equal  subintervals,  only  now  we  require  that  the 
number  of  subintervals  be  even,  which  we  will  write  as  2n.  Thus,  we  take 

b  —  a 

Ax  =  ,     a  =  xo  <  X\  <  . . .  <  X2n  =  b,  where  x,-  =  a  +  /  Ax. 

2n 

Then  the  Simpson's  rule  approximation  Sin  is 

Ax 

Sm  =  —  If  (a)  +  4/(xi)  +  2/(x2)  +  4/(x3)  +  •  ■  ■  +  2/(x2„_2) 
+  4/(x2n_0  +  /(&)] 

n  —  1  n 

f(a)  +  2  £  /(x2()  +  4  £  /(x2l_.)  +  f(b) 


Ax 


(2) 


EXAMPLE  2  We  approximate  the  value  of  f0  sin(x2)  dx  of  Example  1  using 
Simpson's  rule  with  n  =  2  (i.e.,  four  subintervals).  As  before,  we  have 

1  -  0 

Ax  =   =  0.25, 

4 

so  (2)  gives 

■!  o  25 

sin  (x2)  dx  R»  54  =  —  [sin  0  +  4  sin  (0.252)  +  2  sin  (0.52)  +  4  sin  (0.752) 


/ 

Jo 


+  sin  1] 
0.25 


3 

0.25 


[0  +  0.24984  +  0.49481  +  2.13321  +  0.84147] 
[3.71933]  =  0.30994. 


3 

Note  that  this  value  is  in  the  range  predicted  by  the  trapezoidal  rule.  In  fact,  it  is 
a  more  accurate  approximation  to  the  value  of  the  definite  integral.  ♦ 

Simpson's  rule  is  obtained  by  approximating  the  function  /  with  n  quadratic 
functions  that  pass  through  triples  of  points  (x2,_2,  /(x2(_2)),  (x2;_i,  /fe-i)), 
(x2i ,  f(xii))  for  i  =  1 , . . . ,  n.  As  with  the  trapezoidal  rule,  we  have  the  following 
summary  result. 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


THEOREM  7.2    Given  a  function  /  that  is  integrable  on  [a,  b],  we  have 

b 

f(x)dx  =  S2n  +  E2n, 
where 

n—  1  n 

f(a)  +  2  £  /(xa)  +  4  J]  /(x2i_0  +  /(*)  , 
i=i  i=i 

and  Ein  denotes  the  error  involved  in  using  S^i  to  approximate  the  value  of  the 
definite  integral.  In  addition,  if  /  is  of  class  C4  on  [a,  b],  then  there  is  some 
number  £  in  (a ,  b)  such  that 

E,  -  -b~a(Ax)4  f^>(f)  -  - (fc  -  a)  f(4)m 
2"  "      180  {     '  1    (;)  ~     2880/24  1  {;)- 


L 


Sin  = 


Ax 


EXAMPLE  3  Consider  (x3  +  3x2)  dx.  We  compare  the  trapezoidal  rule 
and  Simpson's  rule  approximations  with  4  subintervals.  We  thus  have  Ax  = 
(3  -  l)/4  =  0.5  and 

0.5 

T4  =  —  [4  +  2(10.125  +  20  +  34.375)  +  54]  =  46.75; 
0.5 

S4  =  —  [4  +  4(10.125)  +  2(20)  +  4(34.375)  +  54]  =  46. 

Note  that 

(x3  +  3x2)  dx  =  (I*4  +  x3) \\  =  (f  +  27)  -  (i  +  1)  =  46, 


so  Simpson's  rule  agrees  with  the  exact  answer.  This  should  not  be  a  surprise, 
since  for  f(x)  =  x3  +  3x2,  we  have  that  f{4\x)  is  identically  zero,  which  means 
that  the  error  term  £4  for  Simpson's  rule  must  be  zero.  ♦ 

Approximating  Double  Integrals  over  Rectangles   


Now  let  /  be  a  function  of  two  variables  that  is  continuous  on  the  rectangle 
R  =  [a,  b]  x  [c,  d]  in  R2.  We  adapt  the  previous  ideas  to  provide  methods  for 
approximating  the  value  of  the  double  integral  JjRfdA. 

Because  we  assume  that  /  is  continuous  on  R,  Fubini's  theorem  applies  to 

give 

/ /  f(x,y)dA=  f  f  f(x,y)dydx.  (3) 

J  J R  J  a  Jc 

Next  we  partition  the  x-interval  [a,  b]  into  m  equal  subintervals.  Thus, 
b  —  a 

Ax  =   ,     a  =  xo  <  x\  <  . . .  <  xm  =  b,  where  x, ■  =  a  +  /  Ax. 

m 

Similarly,  we  partition  the  y-interval  [c,  d]  into  n  equal  subintervals;  hence, 
d  —  c 

Ay  =   ,     c  =  yo  <  v\  <  . . .  <  y„  =  d,  where  y;-  =  c  +  j  Ay. 

n 

In  the  inner  integral  fd  f(x,  y)  dy  of  the  iterated  integral  in  (3),  the  variable  x  is 
held  constant.  Therefore,  we  may  approximate  this  integral  using  the  trapezoidal 


Chapter  5  i  Multiple  Integration 


rule  of  Theorem  7.1.  We  obtain 


f{x,y)dy 


Ay 


n-l 


/(ac,c)  +  2^)/(x,y;)  +  /(x,d) 

7  =  1 


Next,  integrate  each  function  of  x  appearing  on  the  right  side,  so  that 

-b  pel 


J  a    J  c 


Ay 
2 


y)dydx  & 

r-b  n  —  1    pb  r-b 

J    f{x,  c)dx  +  2       I    fix^yj)dx+  I  fix,d)dx 

J  a  ,-_  i  J  a  J  a 


(4) 


Now  use  the  trapezoidal  rule  again  on  each  integral  appearing  on  the  right.  This 
means  that,  for  j  =  0, . . . ,  n, 


f 

J  (I 


fix,  yj)dx 


Ax 


f(atyj)  +  2j2f(Xi,yj)  +  m  yj) 


(5) 


Putting  (4)  and  (5)  together,  we  obtain 


p  b  pel 

/   /  /(*, 

J  a    J c 


y)  dy  dx 


Ay  Ax 


m— 1 


n-l 


Ax 


fia,c)  +  2  YJfi.xi,c)  + fib,  c) 

m—  1 

f(a.y,)  •  2j./|v,.v/)  •  /(/>.  V;) 


+ 


7=1 

Ax 


i=i 


fia,d)  +  2YJf(xi,d)  +  fib,d) 


Therefore,  the  trapezoidal  rule  approximation  T„hn  to  ffRf  dA  is 


T 

1  m.n 


Ax  Ay 


-l 


fia,c)  +  2j2f(xi,c)  +  fQ>,c) 


(  =  1 

n— 1 m  — 1 


n— 1  n— 1  m— 1  ft— 1 

+  2  J]  /(a,  y,)  +  4££  /(x, ,  yj)  +  2  J]  fib,  yj)  (6) 

7  =  1  7  =  1  «=1  7  =  1 


1-1 


+  fia,d)  +  2YJf{xi,d)  +  fib,d) 


i=i 


The  expression  appearing  in  (6)  is  not  too  memorable  as  it  stands.  However, 
we  may  interpret  it  as  follows: 


Ax  Ay  -r-^  ^-^ 

Tm,„  =  ——    2^  yj)> 

j=0  1=0 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


where 


1  if  (xi ,  yj)  is  one  of  the  four  vertices  of  R  ; 

2  if  (xi ,  yj)  is  a  point  on  an  edge  of  R,  but  not  a  vertex; 
4     if  (x, ,  yj)  is  a  point  in  the  interior  of  R. 


EXAMPLE  4  We  approximate  the  value  of  L  jx  {xy  +  3x)dydx  with  7^2. 
Thus,  the  ^-interval  [3,  6]  is  partitioned  using  Ajc  =  (6  —  3)/3  =  1  and  the  y- 
interval  [1,  2]  is  partitioned  using  Ay  =  (2  —  l)/2  =  0.5.  See  Figure  5.126  for 
the  rectangle  [3,6]  x  [1,2]  with  partition  points  marked. 
Hence,  we  have 


(xy  +  3x)  dy  dx  « 
1(0.5) 


[12  +  24  +  30+  15 


1 


+  2(16  +  20  +  27  +  20  +  25  +  13.5)  +  4(18  +  22.5)] 
-(486)  =  60.75. 


y 

2  -- 
1.5 
1  + 


15 

13.5, 
12 


20 

25 

30 

18 

22.5 

27 

• 

• 

16 

20 

24 

H  1  1  h 


H  h 


Figure  5.126  The  rectangle 
[3,  6]  x  [1,  2]  of  Example  4  with 
partition  points  marked  with  the 
values  of  f(x,  y)  =  xy  +  3x. 


For  comparison,  we  calculate  the  iterated  integral  directly: 

2 

dx 

y=l 


f6  9  9  2 

=  /    -x  dx  =  -x 
J3   2  4 

Note  that  T3  2  gives  the  exact  answer  in  this  case.  ♦ 

In  the  derivation  above  of  the  trapezoidal  rule  we  assumed  that  the  func- 
tion /  was  continuous.  Nonetheless,  we  may  use  formula  (6)  to  approximate 
ffRf  dA  whenever  /  is  integrable  on  R.  However,  in  order  to  make  estimates 
of  the  accuracy  of  trapezoidal  approximations,  we  must  assume  more  about  /,  as 
the  following  result  (whose  proof  involves  use  of  the  intermediate  value  theorem 
and  the  mean  value  theorem  for  integrals)  indicates. 


if; 


(xy  +  3x)dy  dx 


=  f( 


xy- 
2 


3xy 


243 


=  60.75. 


394       Chapter  5  |  Multiple  Integration 


THEOREM  7.3  (TRAPEZOIDAL  RULE  FOR  DOUBLE  INTEGRALS  OVER  RECTANGLES) 

Given  a  function  /  that  is  integrable  on  the  rectangle  R  =  [a ,  b]  x  [c,  d],  we  have 


//. 


fix,  y)dA  =  Tm%n  +  Em>n, 

R 

where  T„u„  is  given  by  (6)  and  Emj,  denotes  the  error  involved  in  using  T„un  to 
approximate  the  value  of  the  double  integral.  Moreover,  if  /  is  of  class  C2  on  R, 
then  there  exist  points  (fi ,  rj\)  and  (£2,  Vi)  in  R  such  that 

Em,n  =  _(fc-°X^-c)  [(A*)2/^,  ,x)  +  (Av)2/yyfe,  ift)]  • 


EXAMPLE  5  For  /(*,  y)  =  xy  +  3x,  we  have  that  d2f/dx2  and  d2f/dy2  are 
both  identically  zero.  Hence,  Theorem  7.3  implies  that  the  trapezoidal  rule  ap- 
proximation is  exact,  as  we  already  observed  in  Example  4.  ♦ 

In  a  manner  entirely  analogous  to  our  derivation  of  the  trapezoidal  rule,  we 
may  also  produce  a  Simpson's  rule  approximation  to  the  value  of  the  double 
integral  ffR  f  dA.  As  in  the  case  of  Simpson's  rule  for  approximating  single- 
variable  definite  integrals,  we  partition  both  the  x-  and  y-intervals  into  even 
numbers  of  equal  subintervals.  Thus,  we  take 

b  —  a 

Ax  =   ,     a  =  xq  <  x\  <  . . .  <  X2m  =  b,  where  x,-  =  a  +  /  Ax, 

2m 

and 

d  —  c 

Ay  =   ,     c  =  yo  <  y\  <  . . .  <  ym  =  d,  where  y7-  =  c  +  j  Ay. 

2n 

The  resulting  Simpson's  rule  approximation  S2m,2n  is  given  by  the  expression 


Ax  Ay 


m— 1  m 

fia,  c)  +  2j2  f(*2i  ,c)  +  4J2  f(.X2i-u  c)  +  fib,  c) 

(=1  (=1 

—  1  n—  1  m— 1  n— 1  m 


+ 2   f(a,  yij) f(*»>  *y) + 8  E  E 

;=1  7=1  i=l  ;=1  t=l 

n-1 

+  2J2fib,y2j)  (7) 

7  =  1 

+        fia,  yij-x)  +  /(**.  yy-i) 

7=1  ;=1  i=l 

n      m  n 

+ 16  E  E  /c^-i.  ^-i) + 4  E  ^-0 

7=1  i=l  7=1 

m — 1  m 

+  /(fl,  J)  +  2  J]  /(x2„  rf)  +  4£  f(x2i-i,d)  +  fib,  d) 
/=!  i=i 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


Just  as  in  the  case  of  the  trapezoidal  rule  for  double  integrals  over  rectangles,  the 
intermediate  value  theorem  and  the  mean  value  theorem  for  integrals  provide  the 
following  result. 


THEOREM  7.4  (Simpson's  rule  for  double  integrals  over  rectangles) 
Let  /  be  a  function  that  is  integrable  on  the  rectangle  R  =  [a,  b]  x  [c,  d].  Then 


f(x,  y)dA  =  S2,„,2n  +  ^2m,2n> 


where  S2m,2n  is  given  by  (7)  and  E2m,2„  denotes  the  error  involved  in  using  S%m^n 
to  approximate  the  value  of  the  double  integral.  Moreover,  if  /  is  of  class  C4  on 
R,  then  there  exist  points  (£i ,  r]\)  and  (£2,  m)  in  R  such  that 

E2m,2n  =  ~(b~a^  KAX)4 +  (Ay)Vvvvv(£2,  m)]  ■ 


EXAMPLE  6  We  compare  approximations  to  the  value  of  /0°'5  /„'  ex+y  dydx 
using  both  the  trapezoidal  rule  and  Simpson's  rule  with  four  subintervals  in 
each  of  the  x-  and  y-intervals.  Thus,  the  x-interval  [0,  0.5]  is  partitioned  using 
Ax  =  (0.5  —  0)/4  =  0.125  and  the  y-interval  [0,  1]  is  partitioned  using  Ay  = 
(1  -0)/4  =  0.25. 

The  trapezoidal  rule  gives 

T4A  =  (Q-125X°-25>  yO+0  +  e0+0.5  +  el+0  +  e0.5+l  +  2  ^0+0.25  +  e0+0.5 
_|_  ^0+0.75  _j_  e0.125+0  _j_  ^0.25+0  _|_  ^0.375+0  _|_  ^0.125+1  _|_  ^0.25+1 


_|_  ^0.375+1  _|_  ^0.5+0.25  _j_  e0.5+0.5  _j_  ^0.5+0.75^ 

0.125+0.25   |    „0.125+0. 5  _|_  g0. 125+0.75  _|_  g0.25+0.25 
_|_  g0.375+0.5  +  ^0.375+0.75^ j 


+  4(e" 

,  ^0. 25+0. 75   ,  g0.37 


1 

=  (143.608854)  =  1.121944. 

128 

Note  that  d2/dx2  (ex+y)  =  d2/dy2  (ex+y)  =  ex+y .  Theorem  7.3  says  that  there 
exist  points       rji)  and  (£2,  V2)  in  the  rectangle  [0,  0.5]  x  [0,  1]  such  that 

Eaa  =  _(0-5-0)(l  -0)  ^(0_125)2efl+1J1  +  (0.25)V2+"2] . 

On  the  rectangle  [0,  0.5]  x  [0,  1],  the  smallest  possible  value  of  ex+y  is  e0+0  =  1 
and  the  largest  possible  value  is  e0  5+l  =  eLS.  Hence,  we  must  have 

_(Ogl)  j-(0  125)2ei.5  +  (0.25)2^-5]  <  e4A  <  [(0.125)2  +  (0.25)2] 

or 

-0.0145888  <  £44  <  -0.003255. 


Chapter  5  |  Multiple  Integration 


|e0+0.25  +  ^0+0.75  +  g0.125+0  +  ^0.125+1  +  ^0.25+0.5 
g0.375+0   .  g0.375+l    .   g0.5+0.25   ,  ^0.5+0.75-, 


Hence,  the  true  value  of  the  double  integral  lies  between  1.121944  —  0.0145888  = 
1.10735  and  1.121944  -  0.003255  =  1.11869. 
On  the  other  hand,  Simpson's  rule  gives 

54,4  =  (0-125^(0-25)  +  +  eH0  +  ^0.5+1 

+  2  (e°+0-5  +  e°-25+0  +  e°-25+1  +  e°-5+0-5) 
+  4(e 

"  5) 

_|_  g  ^0.125+0.5  _|_  ^0.25+0.25  _|_  g0.25+0.75  _|_  g0.375+0.5^ 

_|_  !  g  ^0.125+0.25  _|_  ^0.125+0.75  _|_  ^0.375+0.25  _|_  ^0.375+0.75^ 

=  —(321.036910)  =  1.114711. 
288 

In  this  case,  we  note  that  d4/dx4  (ex+y)  =  d4/dy4  (ex+y)  =  ex+y,  so,  as  before, 
the  minimum  and  maximum  values  of  these  partial  derivatives  on  [0,  0.5]  x  [0,  1] 
are,  respectively,  1  and  e    .  Therefore,  Theorem  7.4  implies  that 

[(°-125)V-5  +  (0-25)V-5]  <  £4.4  <  [(0.125)4  +  (0.25)4] 

or 

-0.0000516688  <  £4,4  <  -0.0000115289. 

Hence,  the  true  value  of  the  double  integral  lies  between  1.114711  — 
0.0000516688  =  1.11466  and  1.114711  -  0.0000115289  =  1.1147. 
For  comparison,  we  may  calculate  the  iterated  integral  exactly: 

/      /   ex+ydydx=\      (ex+y)f     dx  =  /     (ex+l  -  ex)  dx 
Jo    Jo  Jo  '  Jo 

=  (ex+l  -  ex)\°Q5  =  ehS  -  e0-5  -  e  +  \  ^  1.114686. 

Thus,  we  see  that  Simpson's  rule  gives  a  highly  accurate  approximation  with  a 
very  coarse  partition.  ♦ 


Approximating  Double  Integrals  over  Elementary  Regions 

We  can  modify  the  methods  for  approximating  double  integrals  over  rectangles 
to  approximate  double  integrals  over  more  general  regions.  Suppose  first  that  D 
is  an  elementary  region  of  type  1 ;  that  is, 

D  =  {{x,  y)  6  R2  I  y(x)  <  y  <  S(x),  a  <x  <b}. 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


Then,  if  /  is  continuous  on  D,  Theorem  2.10  tells  us  that 

Iy(x) 


fdA=       /  f(x,y)dydx. 

J  JD  J  a  Jy(x) 


To  approximate  this  iterated  integral  with  a  version  of,  say,  the  trapezoidal  rule, 
we  need  to  partition  the  region  D  in  a  reasonable  way.  We  do  so  as  follows.  First, 
we  partition  the  x -interval  [a,  b]  in  the  usual  way: 

b  —  a 

Ax  =   ,     a  =  xo  <  x\  <  . . .  <  xm  =  b,  where  xl ■  =  a  +  iAx. 

m 

Now,  for  a  fixed x  in  [a,  b],  we  partition  the  corresponding  y-interval  [y(x),  S(x)] 
into  n  equal  subintervals: 

.   .  ,     <5(x)  -  y(x) 
Ay(x)  =  , 


Y(x)  =  yo  <  y\  <  ■  ■  ■  <  yn  =  §(x),  where  yj(x)  =  y(x)  +  JAy(x). 

Note  that  now  Ay  and  the  partition  numbers  yo ,  . . . ,  y„  are  all  functions  of  x .  (See 
Figure  5.127.)  Then,  by  applying  the  trapezoidal  rule  first  to  the  inner  integral 
and  then  the  outer  one,  we  obtain 


pb  pS(x) 
J  a    J  y(x  ) 


fix,  y)dy  dx 


L 


»  Ay(x) 


n-l 


f(x,  yix))  +  2^/(x,  yjix))  +  fix,  Six)) 


dx 


Ax 
~1~ 


Ay  id)  ^— \  Ay(x,  ) 
-——fia,  y(a))  +  2  ^  — —fixi,  y(x,)) 


Ay(fe) 


n-l 


——fia,  yjia))  +  2  ^  — - — y7- (*,■)) 


+  -±2f(b,yjib)) 


+ 


+ 


2 

Ay  (a) 
2 

Ay(&) 


m—  1 


/(a,  5(a))  +  2  J]  /(*,-,  «(*,)) 


398       Chapter  5  |  Multiple  Integration 

The  trapezoidal  rule  approximation  is,  thus, 


T  = 


Ax  Ay  (a) 


4 


-i-i 


f(a,  y(a))  +  2  J]  f(a,  yj(a))  +  f(a,  8(a)) 


AxAy(xi) 


n-l 


2f(Xi ,  y(xt))  +        f(*i ,  (8) 

7  =  1 


+  2/(xi,<5(xi)) 


+ 


AxAy(Z?) 


n-l 


y(6))  +  2  J]  /(ft,  yj(b))  +  /(&,  5(6)) 


EXAMPLE 


4 

3.5 
3 
2.5 

2 


2  2.1  2.2 

Figure  5.1 28  The  partitioned 
region  D  of  Example  7. 


7   We  approximate  /22'2  J"2*  (x3  +  y2)  6?y  dx  by  T2i4.  We  have 
2.2-2 

Ax  =  =  0.1,     so  xo  =  2,  xi  =  2.1,  X2  =  2.2 


and,  therefore,  that 
4-2 

Ay(2)  =  =  0.5, 


so    yo(x0)  =  2,  y1(x0)  =  2.5,  yz(xo)  =  3,  y3(xo)  =  3.5,  y4(x0)  =  4, 


4.2-2.1 

Ay(2.1)  =  =  0.525, 

so    y0(xi)  =  2.1,  yi(xi)  =  2.625,  y2(xi)  =  3.15,  y3(xj)  =  3.675, 
y4Oi)  =  4.2, 

4.4-2.2 

Ay(2.2)  =  =  0.55, 

so    yo(x2)  =  2.2,  yi(x2)  =  2.75,  y2(x2)  =  3.3,  yi(x2)  =  3.85, 
y4(x2)  =  4.4. 

(See  Figure  5.128.)  Thus, 

T2A  =  (°"1)4(Q,5)  [(23  +  22)  +  2  ((23  +  2.52)  +  (23  +  32)  +  (23  +  3.52)) 
+  (23  +  42)] 

+  (0,1)^'525)  [2(2. 13  +  2.12)  +  4(2. 13  +  2.6252)  +  4(2. 13  +  3.152) 

+  4(2. 13  +  3.6752)  +  2(2. 13  +  4.22)] 
+  (0-1^0-55^  [(2.23  +  2.22)  +  2(2.23  +  2.752)  +  2(2.23  +  3.32) 

+  2(2.23  +  3.852)  +  2(2.23  +  4.42)] 
=  0.0125(139)  +  0.02625(156.7755)  +  0.01375(175.934)  =  8.271949375. 


5.7  |  Numerical  Approximations  of  Multiple  Integrals  (optional) 


In  this  case,  the  exact  answer  is 

.2.2    plx  p2.2 


p2.2    p2x  n2.2  n2.2 

J     J     (x3  +  .v2)  dy  dx  =  J      (x3y  +  ±y3)  |^  dx  =  J      (x4  +  |x3)  dx 


=  8.23886. 


I2-2  _  1/T)5 
12  5 


(2.25  -  25)  +  i  (2.24  -  24) 


If  Z)  is  an  elementary  region  of  type  2,  that  is, 

D  =  {(x,  y)  e  R2  |  a(y)  <  x  <  fi(y),  c  <  y  <  d}, 

then  we  may  similarly  approximate  ffDf  dA  by  first  partitioning  the  y-interval 
[c,  d]  using 

d  —  c 

Ay  =   ,     c  =  y0  <  yi  <  . . .  <  y„  =  d,  where  y ;  =  c  +  jAy, 

n 

and  then,  for  a  fixed  y  in  [c,  d],  by  partitioning  the  corresponding  x-interval 
[a(y),  /5(y)]  into  m  equal  subintervals: 

£00  - 


Ajc(y)  = 


77? 


a(y)  =  x0  <  Xi  <  . . .  <  xm  =  /3(y),  where  x,(y)  =  a(y)  +  i  Ax(y). 
In  doing  so,  we  obtain  a  counterpart  formula  to  that  of  (8),  namely, 


T 

±  in 


Ax(c)Ay 


4 

(-1 


|  g  Ax(yy-)Ay 


/(a(c),  c)  +  2      /(*/(c),  c)  +  /(/3(c),  c) 

m-l 

2/(«(y,).yD  •  4j]/(.v,(y,  ).  y;)) 


7  =  1 


1=1 


+  2f(P(yj),yj) 


(9) 


+ 


Ax  (J)  Ay 


m-l 


/(a(d),  d)  +  2j2  f(xi(d),  d)  +  f  (J3(d),  d) 


i=i 


Of  course,  we  may  also  adapt  Simpson's  rule  for  use  in  the  case  of  double 
integrals  over  elementary  regions.  Moreover,  we  may  derive  similar  methods  for 
approximating  triple  integrals  as  well.  In  practice,  however,  other  methods  are 
often  used  that  lend  themselves  to  computer  implementation. 

One  such  alternative  technique  is  known  as  the  Monte  Carlo  method.  It  is 
based  on  a  result  called  the  mean  value  theorem  for  double  integrals:  If  /  is 
a  continuous  function  of  two  variables  and  D  is  a  bounded  and  connected  (i.e., 
one-piece)  region  in  R2,  then  there  is  a  point  P  e  D  such  that 


f(x,  y)dA  =  f(P)  ■  area  of  D. 


400       Chapter  5  |  Multiple  Integration 


5.7  Exercises 


Note  that  it  follows  that 

ffnfdA 
area  ot  D 

so  that,  by  Definition  6.1,  we  have  that  f(P)  =  [f]w%,  the  average  (mean)  value 
of  /  on  D.  The  Monte  Carlo  method  for  approximating  ffDf  dA  is  to  select  n 
points  P\ ,  . . . ,  Pn  in  D  at  random  and  compute  the  average  value  /  of  /  on  just 
these  points: 

Then  /  will  approximate  [/]avg  and,  hence, 


-     (area  of  D) 
fdA~  (area  of  D)f  =  K-  T  f(P,). 

n  '  4 


i=\ 


If  the  area  of  D  is  in  turn  difficult  to  determine  exactly,  it,  too,  may  be  estimated 
by  situating  D  inside  a  rectangle  R  and  selecting  m  points  at  random  in  R.  If  r 
denotes  the  fraction  of  these  points  lying  in  D,  then  area  of  D  «  r  ■  (area  of  R). 

In  an  analogous  manner,  we  may  give  Monte  Carlo  approximations  for  triple 
integrals  of  functions  of  three  variables  defined  over  solid  regions  in  space. 


In  Exercises  1—6,  (a)  use  the  trapezoidal  rule  approximation 
72,3  to  estimate  the  values  of  the  given  integrals,  and  (b)  com- 
pare your  results  with  the  exact  answers. 

*3.1  /-2.1 


l.i  r0.6 


pi. I  nl.\ 

1.  J     J     (x2-6y2)  dydx 

/■3.3  z-3.3 

2.  /      /     xy2  dy  dx 


13  J3 
-2.2  fl.6 


dy  dx 


n 

J      [~Jx  +  <fy)  dy  dx 

/l.l  /.0.6 
J  ex+2ydydx 

6.   /      /      x  cos  y  dy  dx 

J0  Jjt/6 

In  Exercises  7—12,  (a)  use  the  approximation  S2jfrom  Simp- 
son s  rule  to  estimate  the  values  of  the  given  integrals,  and  (b) 
compare  your  results  with  the  exact  answers. 
-OA  /.0.3 


/U.l  />U.J 
/      {y4  -  xy2)  dy  dx 
-o.i  Jo 

pOA  p2 
JO  Jl 


1  +X' 


dy  dx 


Ii  Jo 

Jo  Jn 


?x+2},dvdx 


-7i/4  rn/2 

10.   /       /      sin2x  cos3y  dy  dx 

rr/4 
px/4  rn/4 

11.1       /      sin  (x  +  y)  dy  dx 

Jo  Jo 


1.1  rx/4 


Ii  Jo 


ex  cos  v  dy  dx 


13.  In  Chapter  7  we  will  see  that  the  area  of  the  portion  of 
the  graph  of  f(x,  y)  for  (x,  y)  in  a  region  D  in  R2  is 

given  by  the  double  integral  / fD  J  f  2  +  f2  +  1  dA. 

(a)  Set  up  an  appropriate  iterated  integral  to  compute 
the  surface  area  of  the  portion  of  the  paraboloid 
z  =  4  -  x2  -  3y2  where  (x,  y)  e  [0,  1]  x  [0,  1]. 

(b)  Use  the  trapezoidal  rule  approximation  74  4  to  es- 
timate the  surface  area. 

14.  Concerning  the  iterated  integral 
f\'5  f\.4ln(2x  +  y)  dydx: 

(a)  Calculate  the  trapezoidal  rale  approximation  T^. 

(b)  Use  Theorem  7.3  to  estimate  the  error  in  your  ap- 
proximation in  part  (a). 

(c)  Calculate  the  Simpson's  rule  approximation  52,4. 


True/False  Exercises  for  Chapter  5  401 


(d)  Use  Theorem  7.4  to  estimate  the  error  in  your  ap- 
proximation in  part  (a). 

15.  Without  either  evaluating  or  estimating  the  integral 
fi  4  Jo  s  ln(xy)dy  dx,  which  approximation  is  more 
accurate:  74  4  or  52,2?  Explain  your  answer. 

1 6.  Suppose  that  the  trapezoidal  rule  is  used  to  estimate  the 
value  of  /0°'2  /^qj  exl+2y  dydx.  Determine  the  small- 
est value  of  n  so  that  the  resulting  approximation  T„t„ 
is  accurate  to  within  10~4  of  the  actual  value  of  the 
integral. 

17.  Consider  J0°'3  J0°'4  ex~y  dy  dx. 

(a)  If  the  trapezoidal  rule  approximation  Tn  „  is  used 
to  estimate  the  value  of  this  integral,  what  is 
the  smallest  value  of  n  so  that  the  resulting 
approximation  is  accurate  to  within  10~5  of  the 
actual  value? 

(b)  If  the  Simpson's  rule  approximation  S2,,, i„  is  used 
to  estimate  the  value  of  this  integral,  what  is  the 
smallest  value  of  n  so  that  the  resulting  approx- 
imation is  accurate  to  within  10~5  of  the  actual 
value? 

18.  Concerning  the  iterated  integral  JqJq(3x  + 
5y)dy dx: 

(a)  Calculate  the  trapezoidal  rule  approximation  7/2,2. 

(b)  Compare  your  result  in  part  (a)  with  the  exact 
answer. 

(c)  Use  Theorem  7.3  to  explain  your  results  in  parts 
(a)  and  (b). 

19.  Consider  the  iterated  integral  j\  f^2  x3y3  dy  dx: 

(a)  Calculate  the  Simpson's  rule  approximation  £2,2- 

(b)  Compare  your  result  in  part  (a)  with  the  exact  an- 
swer. 

(c)  Use  Theorem  7.4  to  explain  your  results  in  parts 
(a)  and  (b). 

In  Exercises  20-25,  (a)  use  the  approximation  T3  3  from  the 
trapezoidal  rule  to  estimate  the  values  of  the  given  integrals. 

20.  J  J   (x 3  +  2y2)  dy  dx 


/*jr/4  pcosx 

21 .  /       /       (2x  cos  y  +  sin2  x)  dy  dx 

Jo       J  sinx 
,.0.3  fix 

22.  J     J     (xy  -  x2)  dy  dx 

23.  /        /  dydx 
Jo     Jo      v  1  —  y2 

24.  J  j  sin  x  dx  dy  (Note  the  order  of  integration.) 

/1.6  />2y 
J     In  (xy)dxdy 

26.  In  this  problem,  you  will  develop  another  way  to  think 
about  the  trapezoidal  rule  approximation  given  in  equa- 
tion (6). 

(a)  Let  L  denote  a  general  linear  function  of  two  vari- 
ables, that  is,  L(x,  y)  =  Ax  +  By  +  C,  where  A, 
B,  and  C  are  constants.  Set  R  =  [a,  b]  x  [c,  d]. 
Show  that 

/  /  LdA  =  (area  of  W)(average  of  the  values  of 
•>  JR  L  taken  at  the  four  vertices  of/?). 

(Note  that  this  gives  an  exact  expression  for  the 
double  integral.) 

(b)  Suppose  that  /  is  any  function  of  two  variables  that 
is  integrable  on  R.  Show  that  the  trapezoidal  rule 
approximation  Ti  \  to  ffRf  dA  is 

7^1  =  (area  of  /?)(average  of  the  values  off 
taken  at  the  four  vertices  of  R). 

(c)  Now  let  Ax  =  (b  —  a)/m,  Ay  =  (d  —  c)/n, 
and,  for  i  =  1 , . . . ,  m,  j  =  1, . . . ,  n,  let  /?,•_,•  = 
[Xi-i,  x{\  x  [y,_i,  y,],  where  xt  =  a  +  iAx  and 
y/  =  c  +  j  Ay.  Then  we  have 

///<"  =  ££ //  fdA- 

Use  7iti  to  approximate  each  integral  ffR  fdA 
and  sum  the  results  to  obtain  the  formula  for  T„un 
given  by  equation  (6). 


True/False  Exercises  for  Chapter  5 


1.  Every  rectangle  in  R2  may  be  denoted  [a,  b]  x  [c,  d]. 


f1/2    f2     3  2 

.  I    y  sin  {jix  )  dy  dx 


2.  If  /  is  a  continuous  function  and  f(x,  y)  >  0  on  a 

region  D  in  R2,  then  the  volume  of  the  solid  in  R3  r2  Z"1/2 

under  the  graph  of  the  surface  z  =  f(x,  y)  and  above  =  J  i  Jo     ^  s^n  ^nx  ^ dx 

the  region  D  in  the  xy-plane  is  /  /  f(x,y)dA. 


402       Chapter  5  |  Multiple  Integration 


/     /    3  dy  dx  =  J     J    3  dx  dy 

Jo  Jo  Jo  Jo 


10  Jo 

-2  ry 


5-   /    /    f(x,y)dxdy=  I    I    f(x,  y)dy  dx  for 
Jo  Jo  Jo  Jo 

all  continuous  functions  /. 

6.  0  <  j J  sin(x4  +  y4)dx  dy  <  jt,  where  D  denotes 
the  unit  disk  {(x,  y)  \  x1  +  y2  <  1}. 

7.  f    [  sfx3  +  ydydx  =  [    [  ^x3  +  y  dx  dy. 
Jo  Jo  Jo  Jo 

8.  J        x2ex+y  dydx=(^J  x2ex  dx\(j^  ey  dy^j. 

9.  The  region  in  R2  bounded  by  the  graphs  of  y  =  x3 
and  y  =  ~Jx  is  a  type  3  elementary  region  in  the 
plane. 

10.  The  region  in  R2  bounded  by  the  graphs  of  y  =  sinx, 
y  =  cosx,x  =  u  j  A,  andx  =  5  jr/4  is  a  type  2  elemen- 
tary region  in  the  plane. 

1 1 .  The  region  in  R3  bounded  by  the  graphs  of  y2  —  x2  — 
z2  =  1  and  9x2  +  4z2  =  36  is  a  type  3  elementary 
region  in  space. 

12.  j  j  (y3  +  l)dxdy  gives  the  area  of  the  region  D, 
where  D  =  {(x,  y)  |  (x  -  2)2  +  3y2  <  5}. 

'  // < 

angle  with  vertices  (-1,0),  (1,  0),  (0,  1). 


13.   /  /  (y2  sin  (x3)  +  3)  dx  dy  =  3 /2,  where  D  is  the  tri- 


(x  +  y3+z5)dV  =  0. 

[-2,2]x[-l,l]x[-3,3] 

.   [  [  I  (x  +  z)dV  =  0. 

/[-2,2]x[0,l]x[-l,l] 


SSL 
■SSL 

l l l[-2,2] 
Jo    Jv/4  J- 


x[0,l]x[-l,l] 
4    n^y/2  f2 


(y  +  z)dV  =  0. 
sin  yz  dz  dx  dy  =  0. 


18.  /     /  /   (y  -x2)dzdydx  =  0. 

J-l  J-JT^x2  J-^/l-x2-/ 

19.  If  T(n,  v)  =  (2m  —  v,  u  +  3d),  then  the  area  of  the  im- 
age/) =  T(D*)ofthe  unit  square/)*  =  [0,  1]  x  [0,  1] 
is  7  square  units. 

20.  If  T(m,  v)  =  (v  —  u,3u  +  2v),  then  the  area  of  the  im- 
age D  =  T(D*)  of  the  rectangle  D*  =  [0,  3]  x  [0,  2] 
is  5  square  units. 


21 


77 

JO  Jy 


2  r5-2y 
12 


e2x  y  cos(x  —  3y)  dx  dy 


10  nu/2 


I  f 

JO  Ju 


-5e"  cos  vdvdu. 


22.  If   D    is    the    disk    {(x,  y)  |  x2  +  y2  <  9},  then 
jj  v/9-x2-y2dA  =  36jr. 

/>2jt    f>2  p4 

23.  The  iterated  integral  /      /    /    dzdrdO  represents 

Jo     Jo  J2r 

the  volume  enclosed  by  the  cone  of  height  4  and 
radius  2. 


24.  The  iterated  integral 


ffZS 

J-2  J-V^x2  JO 


{2  +  Jx2  +  y2)dzdy  dx 


is  given  by  an  equivalent  integral  in  cylindrical  coor- 
dinates as 


2m    p2  /.79-r2 


pill     pi  r> 

Jo    Jo  Jo 


(r2  +  2r)dz  dr  d6. 


10    Jo  Jo 
25.  The  iterated  integral 

fV2=^  n^4-x2-y2  

/       /         /  7x2  +  y2  +z2  +5dzdydx 

J-s/2.Jo  J^/x2+y2 


is  given  by  an  equivalent  integral  in  spherical  coordi- 
nates as 

/     /       /    Pv  P2  +  5  sin<pdpd<pd8. 
Jo  Jo  Jo 

26.  The  average  value  of  the  function  /(x,y)  =  x2y 
over  the  semicircular  region  D  =  {(x,  y)  |  0  <  y  < 
V4  —  x2}  is  given  by 

1    rn/2  r2 


or  by 


1   f'  f 

-  /  /  r4sin6»cos26Uri/6» 
x  Jo  Jo 

—  f     [  r4  sin9  cos2  9  drd6. 

X  Jn/2  Jo 


27.  The  center  of  mass  of  a  lamina  represented  by  the  tri- 
angle with  vertices  (2,  0),  (0,  1),  (0,  —1),  and  whose 
density  varies  as  S(x,  y)  =  (x2  +  l)cosy,  has  coordi- 
nates given  by 


fo  fu-2)/2  (*3  +  *) cos  y  dy dx 
r2  rv-*)/ifY2 


y  =  0. 


Jo  !(x-2)/2^x2  +  U  cos  y  dy  dx 

28.  The  centroid  of  a  cone  of  radius  a,  height  h,  with  axis 
the  z-axis  and  vertex  at  (0,  0,  h)  is  (0,  0, 

29.  The  center  of  mass  of  the  solid  cylinder  of  radius  a, 
height  h ,  with  axis  the  z-axis  whose  density  at  any  point 


Miscellaneous  Exercises  for  Chapter  5  403 


30.  The  integral    /  /  /  p5  sin2</5  siri(pdp  dtp  d8  repre- 
J  J  Jw 


varies  as  ed ,  where  d  is  the  square  of  the  distance  from 
the  point  to  the  z-axis  is  (0,  0,  z),  where 

f"  f''  zrerl  dzdrdO 

—  \h.  W  with  density  z,  expressed  in  spherical  coordinates. 


r2n  ra  rh      .2  ,    ,    ,„  sents  the  moment  of  inertia  about  the  z-axis  of  a  solid 

Jo    Jo  Jo  zre  dzdrdd 


fo2"  So  So  re*  dzdr  de 


Miscellaneous  Exercises  for  Chapter  5 


1.  Let  B  be  the  ball  of  radius  3;  that  is, 

B  =  {(x,y,z)\x2  +  y2+z2  <  9}. 

Without  resorting  to  any  explicit  calculation  of  an  iter- 
ated integral,  determine  the  value  of  the  triple  integral 
fffB(z3  +2)dV  by  using  geometry  and  symmetry 
considerations. 

2.  Let  W  denote  half  of  the  solid  ball  of  radius  2;  that  is, 

W  =  {(x,  y,  z)  |  x1  +  y2  +  z2  <  4,  z  >  0}. 

Without  resorting  to  explicit  calculation  of  an  iterated 
integral,  determine  the  value  of  the  triple  integral 


SSL 


(x3+y-3)dV. 


(Hint:  Use  geometry  and  symmetry.) 

3.  Let  W  be  the  solid  region  in  R3  with  x  >  0  that  is 
bounded  by  the  three  surfaces  z  =  9  —  x2,  z  =  2x2  + 
y2,  and  x  =  0. 

(a)  Set  up,  but  do  not  evaluate,  two  different  (but 
equivalent)  iterated  integrals  that  both  give  the 
vzlueof  fffw  3  dV. 

(b)  Use  a  computer  algebra  system  to  find  the  value 
of  Jffw3dV  and  to  check  for  consistency  in  your 
answers  in  part  (a). 

4.  Suppose  that  /  is  continuous  on  the  rectangle  R  = 
[a,  b]  x  [c,  d].  For  (x,  y)  G  (a,  b)  x  (c,  d),  we  define 


F(x,y)- 


ff 

J  a    J  c 


f(x',  y')dy'dx' 


Use  Fubini's  theorem  to  show  that  d2F/dxdy  = 
d2F/dydx.  This  provides  an  alternative  proof  of  the 
equality  of  mixed  partials.  (Hint:  Write 


where 


F(x,  y) 


*(*',  y) 


f  g(x', 

J  a 


y)dx' 


j"  f(x',y')dy'. 


Then  dF/dx  and  d2F/dydx  may  be  calculated  using 
the  fundamental  theorem  of  calculus.  Then  use  Fubini's 
theorem  to  find  dF/dy  and  d2 F/dxdy.) 


5.  Convert  the  following  cylindrical  integral  to  equivalent 
iterated  integrals  in  (a)  Cartesian  coordinates  and  (b) 
spherical  coordinates: 


p2n    p\    p -\ 

Jo    Jo  Jo 


r  dz  dr  d6. 

lo    Jo  Jo 

Evaluate  the  easiest  of  the  three  iterated  integrals. 
6.  The  volume  of  a  solid  is  given  by  the  iterated  integral 


Jo  Jo  J-^/i 


dz  dx  dy. 


(a)  Sketch  the  solid  and  also  describe  it  by  giving  equa- 
tions for  the  surfaces  that  form  its  boundary. 

(b)  Express  the  volume  as  an  iterated  integral  in  cylin- 
drical coordinates.  Determine  the  volume. 

7.  Calculate  the  volume  of  a  cube  having  edge  length  a 
by  integrating  in  cylindrical  coordinates.  (Hint:  Put  the 
center  of  the  cube  at  the  origin.) 

8.  Calculate  the  volume  of  a  cube  having  edge  length  a 
by  integrating  in  spherical  coordinates. 


9.  Determine 


cos 


2y 


X  +  V 


dA, 


where  D  is  the  triangular  region  bounded  by  the  coor- 
dinate axes  and  the  line  x  +  y  =  1 . 


10.  Evaluate 


/>6  />1—  2y 
Jo  J-2v 


y\x  +  2y)2e(x+2y?  dxdy 


by  making  a  suitable  change  of  variables. 

1 1 .  Find  the  area  enclosed  by  the  ellipse  E  given  by  the 
equation 


~  +  ~b~2 


1 


a 


in  the  following  way: 

(a)  First,  write  the  area  as  the  value  of  an  appropriate 
iterated  integral  in  Cartesian  coordinates.  Do  not 
evaluate  this  integral. 

(b)  Next,  scale  the  variables  by  letting  x  =  ax, 
y  =  by.  To  what  region  E*  in  the  iy-plane  does 


404       Chapter  5  ,  Multiple  Integration 


the  ellipse  E  correspond?  Rewrite  the  xy-integral 
in  part  (a)  as  an  xy-integral. 

(c)  Finally,  use  polar  coordinates  to  transform  the  xy- 
integral  and  thereby  show  that  the  area  inside  the 
original  ellipse  is  nab. 

1 2.  This  problem  concerns  the  rotated  ellipse  E  with  equa- 
tion ttx1  +  I4xy  +  10y2  =  9. 
(a)  Let  u  =  2x  —  y,  v  =  x  +  y  and  rewrite  the  equa- 
tion for  E  in  the  form 


+ 


b- 


l, 


where  a  and  b  are  positive  constants  to  be 
determined. 

(b)  Use  an  appropriate  change  of  variables  and  the  re- 
sult of  part  (c)  of  Exercise  1 1  to  find  the  area  en- 
closed by  E . 

13.  Consider  the  ellipse  E  with  equation 
5x2  +  6xy  +  5y2  =  4.  Let  u  =  x  —  y,  v  =  x  +  y  and 
follow  the  steps  of  Exercise  12. 

1 4.  Imitate  the  techniques  of  Exercise  1 1  to  find  the  volume 
enclosed  by  the  ellipsoid  E  given  by  the  equation 


2  2  2 

x       y  Z 

 h  1  

a2      b2  c2 


1. 


15.  Evaluate 


xy 


d  y  —  x 


dA, 


where  D  is  the  region  in  the  first  quadrant  bounded 


by  the  hyperbolas  x 


y 


1,  x  —  y  =  4  and 


the  ellipses  x2/4  +  y2  =  1,  x2/\6  +  y2/4  =  1.  (Hint: 
Sketch  the  region  D,  and  use  it  to  make  an  appropriate 
change  of  variables.) 


16.  Evaluate 


(x2  +  y  V 


dA, 


where  D  is  the  region  in  the  first  quadrant  bounded 
by  the  hyperbolas  x2  —  y2  =  1,  x2  —  y2  =  9,  xy  =  1, 
xy  =  4. 


17.  Evaluate 


dA, 


where  D  is  the  region  bounded  by  xy  =  1,  xy  =  4, 

y  =  l,  y  =  2. 

1 8.  (a)  Generalizing  the  notions  of  double  and  triple  inte- 
grals, develop  a  definition  of  the  "quadruple  inte- 
gral" // ffw  fdVofa  function  f(x,  y,  z,  w)  over 
a  four-dimensional  region  W  in  R4. 


mi 


(b)  Use  your  definition  in  part  (a)  to  calculate 
(x  +  2y  +  3z-4w)dV, 

r 

where  W  is  the  four-dimensional  box 

W  =  {(x,  y,  z,  w)  |  0  <  x  <  2,  -1  <  y  <  3, 
0  <  z  <  4,  -2  <  w  <  2}. 

19.  (a)  Set  up,  but  do  not  evaluate,  a  quadruple  iterated  in- 

tegral that  computes  the  four-dimensional  volume 
of  the  four-dimensional  ball  of  radius  a : 

B  =  {(x,  y,  z,  w)  |  x2  +  y2  +  z2  +  w2  <  a2}. 

(b)  Use  a  computer  algebra  system  to  give  a  formula 
for  the  volume. 

(c)  Use  a  computer  algebra  system  to  give  for- 
mulas for  the  n-dimensional  volume  of  the 
M-dimensional  ball 

B  =  {{xx,x2,  . . . ,  Xn)  I  x\  +  x\  H  Yx2n<  a2} 

in  the  cases  where  n  =  5,  6.  Is  there  any  pattern  to 
your  answers? 

In  Exercises  20—23,  you  will  give  a  general  expression  for  the 
n-dimensional  volume  of  the  n-dimensional  ball 

B  =  {(xx,x2,...,xn)  |  x2+x2  +  ---x2  <  a2}. 

Let  V„(a)  denote  this  n-dimensional  volume.  Let  C„  denote  the 
n-dimensional  volume  of  the  unit  ball  U;  that  is,  C„  =  V„(l). 

20.  By  scaling  the  variables  x\,x2,  ■  ■  ■  ,xn  and  using  a 
change  of  variables  formula  analogous  to  those  in  The- 
orems 5.3  and  5.5,  show  that  V„(a)  =  Cnan. 

21.  (a)  Consider  points  in  B  of  the  form  (jci,  x2,  0,  0). 

Show  that  the  coordinates  x\,  x2  describe  points 
(in  R2)  lying  in  the  disk  of  radius  a. 

(b)  In  R",  let  the  polar  coordinates  r  and  9  replace 
the  Cartesian  coordinates  X\  and  x2.  Argue  that 
the  points  in  B  lying  over  a  particular  point  (r,  6) 
in  the  disk  described  in  part  (a)  must  fill  out  an 
(n  —  2)-dimensional  ball  of  radius  V 'a2  —  r2. 

(c)  Use  part  (b)  to  show  that  the  w-dimensional  volume 
of  B  is  given  by 

f2jT 

■de. 


f  I  Vn.2(Va2-r2)rdr, 
Jo  Jo 


22.  Use  the  previous  two  exercises  to  establish  the  recur- 
sive formula 

(2jta2\ 
Vn(a)  =    Vn-2{a). 


23.  (a)  Show  that  V\{a)  =  2a  and  V2(a)  =  it  a2.  (These 
are  familiar  facts.) 


Miscellaneous  Exercises  for  Chapter  5  405 


(b)  Use  part  (a)  and  the  previous  problem  to  show  that 


Vn(a)  : 


7T"'2 

(n/2)! 

2(n+l)/2jr(«-l)/2 


if  n  is  even 
a"     if  n  is  odd 


where  the  double  factorial  n\\  =  n(n  —  2)(«  — 
4)  •  •  •  3  •  1  (i.e.,  the  product  of  all  odd  integers 
from  1  to  n). 

24.  A  spherical  shell  with  inner  radius  3  cm  and  outer 
radius  4  cm  has  a  mass  density  that  varies  as 
0.12c?2  g/cm3,  where  d  denotes  the  distance  (in  cen- 
timeters) from  a  point  in  the  shell  to  the  center  of  the 
shell. 

(a)  Determine  the  total  mass  of  the  shell. 

(b)  Will  the  shell  float  in  water?  (Note:  The  density 
of  water  is  1  g/cm3.  To  answer  this  question,  you 
need  to  determine  the  average  density  of  the  shell.) 

(c)  Suppose  that  the  shell  has  a  small  hole  so  that  the 
core  of  the  shell  fills  with  water.  Now  will  it  float? 

25.  A  dome  is  shaped  as  a  hemisphere.  If  a  pole  whose 
length  is  the  average  height  of  the  dome  is  to  be  in- 
stalled inside  the  dome  in  an  upright  position,  where 
on  the  floor  can  it  be  located? 

26.  Let  /  be  continuous  on  R  =  [a,  b]  x  [c,  d].  In  this 
problem,  you  will  establish  Leibniz's  rule  for  "differ- 
entiating under  the  integral  sign": 


d 

dy 


f 


f(x,  y)dx 


J  a 


(x,  y)  dx. 


(a)  Let  G(y')  =  fy(x,  y')dx.  For  c  <  y  <  d,  use 
the  fundamental  theorem  of  calculus  to  compute 
d/dy  P  G(y')dy'  and,  therefore, 


d 
dy 


-  /     /    fy{x,y')dxdy' . 

y  Jc    J  a 


(b)  Use  Fubini's  theorem  and  part  (a)  to  establish 
Leibniz's  rule. 

27.  The  function  f(x,  y)  =  1/ ^fxy  is  unbounded  when  ei- 
ther x  or  y  is  zero.  Thus,  if  D  =  [0,  1]  x  [0,  1],  we 

say  that  j  j    —  dA  is  an  improper  double  inte- 


>xy 

gral,  analogous  to  the  one-variable  improper  integral 
f1  1 

/  — —  dx.  Improper  multiple  integrals  of  this  type 
Jo  V* 

may  be  evaluated  using  an  appropriate  limiting  pro- 
cess. In  this  problem,  you  will  determine  the  value  of 

j  j    —  d  A  in  the  following  manner: 


Id  Jxy 


(a)  For  0  <  e  <  1/2,  0<<5<l/2,  let  D,j  = 
[e,  1  -  e]  x  [8,  1  -  8].  (Note  that  D€.s  c  D.) 
Calculate 


7(6,  8) : 


1 


dA. 


(b)  Evaluate     lim     I(e,  8).  You  should  obtain  a  fi- 

M)-K0,0) 

nite  value,  which  may  be  taken  to  be  the  value  of 
j  j    —  dA,  since  D(  s  "fills  out"  D  as  (e,  8) 


Id  -jxy 

(0,  0).  (We  say  that  in  this  case  the  improper  inte- 
gral converges.) 

28.  Imitate  the  techniques  of  Exercise  27  to  determine  if 
the  improper  double  integral 


a 


i 


dA 


[o,i]x[o,i]  x  +  y 

converges  and,  if  it  does,  find  its  value.  (Hint:  You  will 
need  to  determine  lim„_!.o+  u  In  u.) 

29.  Imitate  the  techniques  of  Exercise  27  to  determine  if 
the  improper  double  integral 


I I\o, 


dA 


/[o,i]x[o,i]  y 

converges  and,  if  it  does,  find  its  value. 

30.  Calculate  / fD  In  y ' x1  +  y2  dA,  where  D  is  the  unit 

disk  x2  +  y2  <  1 .  Note  that  the  integrand  is  not  de- 
fined at  the  origin,  so  this  is  an  example  of  an  improper 
double  integral.  Nonetheless,  you  can  find  its  value  by 
integrating  over  the  annular  region 


{(x,y)  |  €  <x2  +  y2  <  1} 


and  taking  appropriate  limits. 

31 .  Find  the  value  of  the  improper  triple  integral 

In  ^x2  +  y2  +  z2dV, 

where  B  is  the  solid  ball  x2  +  y2  +  z2  <  1 .  (See  Exer- 
cise 30.) 

32.  If  D  is  an  unbounded  region  in  R2,  the  integral 
ffD  f(x,  y)dA  is  another  type  of  improper  double 
integral,  analogous  to  one-variable  improper  integrals 
such  as  /fl°°  f(x)dx,  f(x)dx,  or  f™^  f(x)  dx.  In 
this  problem,  you  will  determine  the  value  of 


IL 


i 


d  x2y3 


dA, 


where  D  =  {(x,  y)  |  x  >  1,  y  >  1}  using  a  limiting 
process. 


406       Chapter  5  |  Multiple  Integration 


(a)  For  a  >  1,  b  >  1,  let  Da  j,  =  [1,  a]  x  [1,  b].  Com- 


pute 


I(a,b) 


dA. 


(b)  Evaluate      lim      I  (a,  b).  You  should  obtain  a 

(a,b)— >(oo,oo) 

finite  value,  which  may  be  taken  to  be  the  value 

of  f  f  — r — -  dA.  (In  such  a  case,  we  say  that  the 

J  Jd  *2r 
improper  integral  converges.) 

33.  Let  D  =  l(x,  y)  |  x  >  1,  y  >  1}.  For  what  values  of  p 
and  q  does  the  improper  integral 

1 


11 


xPyi 


dA 


converge?  For  those  values  of  p  and  q  for  which  the 
integral  converges,  what  is  the  value  of  the  integral? 

34.  This  problem  concerns  the  improper  integral 

n  +  x2  +  y2y]dA, 


R 

where  p  is  a  constant. 

(a)  Determine  if  the  integral  converges  when  p  =  —2 
by  integrating  over  the  disk  Da  =  {(x,  y)  \  x2  + 
y2  <  a2}  and  then  letting  a  — >  oo. 

(b)  Determine  for  what  values  of  p  the  integral 
//jj2(1  +  x2  +  y2)p  dA  converges.  What  is  the 
value  of  the  integral  when  it  converges? 


35.  Determine  if 


SSL 


1 


R3  (1  +  x2  +  y2  +  z2fl2 


dV 


converges  by  integrating  over  the  ball  Ba  =  [(x,  y,  z) 
x2  +  y2  +  z2  <  a2}  and  then  letting  a  — >  oo. 


36.  Show  that 


SSL 


-Jx^+Z1  dy 


converges  by  determining  its  value. 

37.  In  this  problem,  you  will  find  the  value  of  the  one- 
variable  improper  integral  f™  e~x  dx  by  using  two- 
variable  improper  integrals. 

(a)  First  argue  that  J°°  e~xl  dx  converges.  (Hint: 

Note  that  e~xl  <  l/x2  for  all  x;  compare 
integrals.) 

(b)  Let  I  denote  the  value  of       e~xl  dx.  Show  that 


/OO  rC 
-oo  J  —  c 


e  x   y  dx  dy 


/r2 

(c)  Let  Da  denote  the  disk  x2  +  y2  <  a2.  Evaluate 

ffDae-*2-y>dA. 


SL<-' 


dA. 


(d)  Compute  I2  as  lima^oo 

(e)  Now  find       e~xl  dx. 


'Da 


dA. 


Exercises  38-46  involve  the  notion  of  probability  densities.  A 
probability  density  function  of  a  single  variable  is  any  func- 
tion f(x)such  that  f(x)  >  0  for  all  x  G  R,  and  f(x)  =  1. 
Given  such  a  density  function,  the  probability  that  a  randomly 
selected  number  x  falls  between  the  values  a  and  b  is 


Prob(a  <  x  <  b) 


f 

J  a 


f{x)dx. 


38.  (a)  Check  that  f(x)  =  e  2'*'  is  a  probability  density 
function. 

(b)  Egbert  turns  on  the  stove  in  a  random  manner  to 
heat  cooking  oil  to  fry  chicken.  If  the  probability 
density  of  the  temperature  x  of  the  oil  is  given  by 
f(x)  =  ie-lx-30°l,  what  is  the  probability  that  the 
oil  has  a  temperature  between  250  °F  and  350 °F? 

A  joint  probability  density  function  for  two  random  variables 
x  and  y  is  a  function  fix,  y)  such  that 

(i)  f(x,  y)  >  0  for  all  (x,  y)  e  R2,  and 

(ii)  //R2  /(*,  y)  dA  =  JZo  f-oo       y)dx  dy=\. 
If  f  is  such  a  probability  density  and  D  is  a  region  in  R2,  then 
the  probability  that  a  randomly  chosen  point  (x,  y)  lies  in  D  is 


Prob((jc,  y)  e  D)  = 

39.  (a)  Show  that  the  function 

I2x  +  y 
140 
0 


f(x,y)dA. 


if  0  <  x  <  5,  0  <  v  <  4 


otherwise 
is  a  joint  probability  density  function, 
(b)  Find  the  probability  that  x  <  1  and  y  <  1 . 
40.  (a)  Show  that  the  function 


f(^y) 


ye 


0 


if  x  >  0,  y  >  0 
otherwise 


is  a  joint  probability  density  function, 
(b)  What  is  the  probability  that  x  +  y  <  2? 

41 .  If  a  and  b  are  fixed  positive  constants,  what  value  of  C 
will  make  the  function  f(x,  y)  =  Ce~a^~h^  a  joint 
probability  density  function? 

42.  Let  a  and  b  be  fixed  nonnegative  constants,  not  both 
zero.  For  what  value  of  C  is 


fi^y) 


\C(ax  +  by)    if  0  <  x  <  1,  0  <  y  <  1 
0  otherwise 
a  joint  probability  density  function? 


Miscellaneous  Exercises  for  Chapter  5  407 


43.  Let  a  and  b  be  fixed  positive  constants  and,  for  a  given 
constant  C,  consider  the  function 


f(x,y) 


Cxy    if  0  <  x  <  a,  0  <  y  <b 
0  otherwise 


(a)  For  what  value  of  C  is  /  a  joint  probability  density 
function? 

(b)  Using  the  value  of  C  that  you  found  in  part  (a), 
what  is  the  probability  that  bx  —  ay  >  01 

44.  The  research  team  for  Vertigo  Amusement  Park  deter- 
mines that  the  length  of  time  x  (in  minutes)  a  customer 
spends  waiting  to  participate  in  the  new  Drown  Town 
water  ride,  and  the  length  of  time  y  actually  spent  in  the 
ride,  are  jointly  distributed  according  to  the  probability 
density  function 

-j:/50-v/5 

fix,  y) 


250" 


0 


if*  >  0,y  >  0 
otherwise 


Find  the  probability  that  a  customer  spends  at  most 
an  hour  involved  with  the  ride  (both  waiting  and 
participating). 

45.  Suppose  that  you  randomly  shoot  arrows  at  a  circular 
target  so  that  the  distribution  of  your  arrows  is  given 
by  the  probability  density  function 


fix,  y) 


1 


where  x  and  y  are  measured  in  feet.  In  the  center  of 
the  target,  there  is  a  bull's-eye  that  measures  1  ft  in 
diameter.  (See  Figure  5.129.)  What  is  the  probability 
of  your  hitting  the  bull's-eye? 


Figure  5.129  The  circular 
target  of  Exercise  45.  The 
shaded  region  is  the 
bull's-eye. 

46.  If  x  is  a  random  variable  with  probability  density  func- 
tion f(x)  and  y  is  a  random  variable  with  probability 
density  function  g(y),  then  we  say  that  x  and  y  are 
independent  random  variables  if  their  joint  density 
function  is  the  product  of  their  individual  density  func- 
tions, that  is,  if 

F(x,y)  =  f(x)g(y). 


Suppose  that  an  electrical  circuit  is  designed  with  two 
identical  components  whose  time  x  to  failure  (mea- 
sured in  hours)  is  given  by  an  exponential  probability 
density  function 


fix)  = 


0  if  x  <  0 

_J_g-Jt/2000  jf  >  0 ' 
2000e 


Assuming  that  the  components  fail  independently, 
what  is  the  probability  that  they  both  fail  within  2000 
hours? 

47.  Let  IL  denote  the  moment  of  inertia  of  a  solid  W  (of 
density  S(x,  y,  z))  about  the  line  L.  (See  formula  (8)  in 
§5.6.)  Let  L  be  the  line  parallel  to  L  that  passes  through 
the  center  of  mass  of  W.  Then,  if  M  denotes  the  mass 
of  W  and  h  the  distance  between  L  and  L,  prove  the 
parallel  axis  theorem: 


48. 


49. 


If 


Mh1 


(Hint:  Without  loss  of  generality,  you  can  arrange 
things  so  that  L  is  the  z-axis.) 

Let  F  be  a  function  of  one  variable  that  is  continuous  on 
the  interval  [a,  b].  Then  the  function  f(x,  y)  =  F{x) 
(i.e.,  the  function  F  considered  as  a  function  of  two 
variables)  is  continuous  on  the  rectangle  R  =  [a,  b]  x 
[c,  d]. 

(a)  Show  that  the  trapezoidal  rule  approximation  Tln  n 

t0  / [r  /(*>  y) dA  is  =id  ~  c)T>n,  where  Tm 
denotes  the  trapezoidal  rule  approximation  to  the 

definite  integral  F(x)dx. 

(b)  Similarly,  show  that  the  Simpson's  rule  approxi- 
mation S2m.2„  to  ffR  f(x,  y)dA  is  S2m,2„  =  (d  - 
c)S2m,  where  S2,„  denotes  the  Simpson's  rule  ap- 
proximation to  the  definite  integral  fb  F(x)  dx. 

Suppose  that  /  is  a  function  of  one  variable  that  is 
continuous  on  [a,  b],  and  g  is  another  function  of  one 
variable  that  is  continuous  on  [c,  d].  Then,  from  Exer- 
cise 40  in  85.2,  we  know  that 


l l[a 


fix)giy)dA 

[a,b]x[c,d] 

f{x)dx\U  giy)dy 


Show  that  this  same  product  property  holds  for 
the  trapezoidal  rule  approximation.  That  is,  if 
Tmi„  denotes  the  trapezoidal  rule  approximation 


t0  IL 


[a,b]x[c,d] 


f(x)g(y)d A,  then 


T,nif)Tn(g), 


where  Tm(f)  denotes  the  trapezoidal  rule  approxima- 
tion to  /*  f(x)dx  and  Tn(g)  the  trapezoidal  rule  ap- 
proximation to  J!!  giy)dy. 


Line  Integrals 


6.1  Scalar  and  Vector  Line 
Integrals 

6.2  Green's  Theorem 

6.3  Conservative  Vector  Fields 

True/False  Exercises  for 
Chapter  6 

Miscellaneous  Exercises  for 
Chapter  6 


x(«)  x(f  \    \S     tek  =  arclength 
*• —  J  1L —  of  segment 

Figure  6.1  The  sum 

Yl"k=\  f(x(tt))^sk  approximates 
the  total  charge  along  an  idealized 
wire  described  by  the  path  x. 


6.1    Scalar  and  Vector  Line  Integrals 

In  this  section,  we  describe  two  methods  of  integrating  along  a  curve  in  space  and 
explore  the  meaning  and  significance  of  our  constructions.  The  main  definitions 
are  stated  first  for  parametrized  paths.  Ultimately,  we  show  that  the  integrals 
denned  are  largely  independent  of  the  parametrization,  that  instead  they  reflect 
essentially  only  the  geometry  of  the  underlying  curve. 

Scalar  Line  Integrals   

To  begin,  we  find  a  way  to  integrate  a  function  (a  scalar  field)  along  a  path. 
Let  x:  [a,  b]  — >  R3  be  a  path  of  class  Cl.  Let  /:  X  c  R3  — >  R  be  a  continuous 
function  whose  domain  X  contains  the  image  of  x  (so  that  the  composite  /(x(f )) 
is  denned).  As  has  been  the  case  with  every  other  integral,  the  scalar  line  integral 
is  a  limit  of  appropriate  Riemann  sums. 
Let 


a  =  to  <  t\  < 


<  h  < 


<t„=b 


be  a  partition  of  [a ,  b] .  Let  tk  be  an  arbitrary  point  in  the  kth  subinterval  [f*_i ,  tk  ] 
of  the  partition.  Then  we  consider  the  sum 


(i) 


k=l 


the 


where  Ask  =  ffk^  ||x'(0||  dt  is  the  length  of  the  &th  segment  of  x  (i.e. 
portion  of  x  denned  for  tk-\  <t<  tk).  If  we  think  of  the  image  of  the  path  x 
as  representing  an  idealized  wire  in  space  and  /(x(^))  as  the  electrical  charge 
density  of  the  wire  at  a  "test  point"  x(f^)  in  the  £th  segment,  then  the  product 
f(x(tk))Ask  approximates  the  charge  contributed  by  the  segment  of  curve,  and 
the  sum  in  ( 1 )  approximates  the  total  charge  of  the  wire.  (See  Figure  6. 1 .)  To  find 
the  actual  charge  on  the  wire,  it  is  reasonable  to  take  a  limit  as  the  curve  segments 
become  smaller;  that  is, 


Total  charge  =     lim     /    f(x(tk  ))Ask 

all  Ait^0  ^ 


all  Alt->0  j' 


(2) 


since  x  is  of  class  C  . 


6.1  |  Scalar  and  Vector  Line  Integrals  409 


The  mean  value  theorem  for  integrals1  tells  us  that  there  is  some  number  f£* 
in  \tk-i,  tk]  such  that 

Ask  =  f    ||x'(?)||  dt  =  (tk  -  tk-i)  \\x'(tk**)\\  =  h'(tk**)\\  Atk- 

(Here  Atk  denotes  tk  —  tk-\.)  Since  ?,*  is  an  arbitrary  point  in  [tk-i,  tk],  we  may 
take  it  to  be  equal  to  .  Therefore,  by  substituting  for  As*  in  equation  (2),  and 
letting  t*k  equal  f|*,  we  have 

n 

Total  charge  =    lim    J]  /W/D)  Aft 
=  fbf(x(t))\\x'(t)\\dt. 

J  a 

This  last  result  prompts  the  following  definition: 


DEFINITION  1 .1    The  scalar  line  integral  of  /  along  the  C1  path  x  is 

f  /(x(f))  ||x'(?)||  dt. 
We  denote  this  integral  fxfds. 


EXAMPLE  1  Let  x:  [0,  2jt]  ->  R3  be  the  helix  x(?)  =  (cos?,  sin?,  ?)  and  let 
/(x,  v,  z)  =  x y  +  z.  We  compute 

ffds  =  J  f(x(t))\\x'(t)\\dt. 


First, 


so  that 


We  also  have 


x'(t)  =  (—  sin?,  cos?,  1), 
s'(0||  =  v7 sin2  ?  +  cos2  f  +  1  =  \/2. 


/(x(/))  =  cos?  sin?  +  f  =  i  sin 2?  +  ? 
from  the  double-angle  formula.  Thus, 


z1  /•  zjt  z1  zjt 

fds=       {\  sin2?  +  ?)  Vldt  =  V2  /    (|  sin2?  +  ?) 

=  V2  (- 1  cos  2?  +  \t>)  |f  =  V2  ((- \  +  2x*)  -(4  +  0)) 


2V27T 


2 


1  Recall  that  this  theorem  says  that  if  F  is  a  continuous  function,  then  there  is  some  number  c  with 
a  <  c  <  b  such  that  /*  F(x)dx  =  (b  -  a)F(c). 


410       Chapter  6  |  Line  Integrals 


Figure  6.2  ApiecewiseC1 
path  x. 


Given  the  discussion  preceding  the  formal  definition  of  the  scalar  line  integral, 
it  is  both  convenient  and  appropriate  to  view  the  notation  fx  f  ds  as  suggesting 
that  the  line  integral  represents  a  sum  of  values  of  /  along  x  times  "infinitesimal" 
pieces  of  arclength  of  x. 

Definition  1 . 1  is  made  only  for  paths  x  in  R3  and  functions  /  defined  on 
domains  in  R3.  Nonetheless,  for  arbitrary  n,  we  may  certainly  use  the  definite 
integral 


f  f(x(t))  \\x'(t)\\  dt, 

J  a 


where  x  is  a  C1  path  in  R"  and  /  is  an  appropriate  function  of  n  variables.  We 
call  this  definite  integral  the  scalar  line  integral  as  well  (and  maintain  the  notation 
J  f  ds)  and  rely  on  the  context  to  make  clear  the  dimensionality  of  the  situation. 
Also,  if  x  is  not  of  class  C1,  but  only  "piecewise  C1"  (meaning  that  x  can  be 
broken  into  a  finite  number  of  segments  that  are  individually  of  class  C1),  then 
we  may  still  define  the  scalar  line  integral  f  f  ds  by  breaking  it  up  in  a  suitable 
manner.  A  similar  technique  must  be  used  if  / (x(?))  is  only  piecewise  continuous. 

EXAMPLE  2    Let  f(x,  y)  =  y  -  x  and  let  x:  [0,  3]  ->  R2  be  the  planar  path 

(2t,  t)  ifO  <  t  <  1 


x(/)  = 


(r+  1,5-4/)   ifl  <t  <3' 


Hence,  x  is  piecewise  C1  (see  Figure  6.2);  the  two  path  segments  defined  for  t  in 
[0,  1]  and  for  /  in  [1,  3]  are  each  of  class  C1.  Thus, 


L 


f  ds 


I  fds+i  f  ds, 


where  xj(/)  =  (It,  t)  for  0  <  t  <  1  and  x2(0  =  (f  +  1,  5  -  4f )  for  1  <  t  <  3. 
Note  that 


|x;(o|  =  VI 


and    fx' (Oil  =  VT7. 


Consequently, 


/  fds=  f  /(x(0)  ||x'(0||  dt  =  f  (t-2t)-V5dt  = 
Jm  Jo  Jo 


V5 


V5 
2 


Also, 


Lfds=L 


/(x2(0)  1 4(0  II  dt 


/I<(5 

17  (At 


■4t)-(t+  l)Wl7dt 


■12717. 


Hence, 


J, 


fds 


V5 
2 


12V17. 


Vector  Line  Integrals  — 

Now  we  see  how  to  integrate  a  vector  field  along  a  path.  Again,  let  x:  [a,  b]  —>  R" 
be  of  class  C1.  (n  will  be  2  or  3  in  the  examples  that  follow.)  Let  F  be  a  vector 


6.1  |  Scalar  and  Vector  Line  Integrals       41  ' 


field  defined  on  a  subset  X  of  R"  such  that  X  contains  the  image  of  x.  Assume 
that  F  varies  continuously  along  x. 


DEFINITION  1 .2  The  vector  line  integral  of  F  along  x:  [a,  b]  R",  de- 
noted /XF  -ds,  is 


f  F-  ds  =  I  F(x(t))-x'(t)dt. 

Jx  J  a 


We  caution  you  to  be  clear  about  notation.  In  the  vector  line  integral  f  F  •  ds, 
the  differential  term  ds  should  be  thought  of  as  a  vector  quantity  (namely,  the 
"differential"  of  position  along  the  path),  whereas  in  the  scalar  line  integral 
/x  f  ds,  the  differential  term  ds  is  a  scalar  quantity  (namely,  the  differential  of 
arclength). 

EXAMPLE  3    Let  F  be  the  radial  vector  field  on  R3  given  by  F  =  xi  +  yj  +  zk 

and  let  x:  [0,  1]  ->  R3  be  the  path  x(t)  =  (t,  3t2,  2t3).  Then  x\t)  =  (1,  6t,  6t2), 
and  so 

j^¥-ds  =  J  F(x(t))-x'(t)dt 

=  f  (ti  +  3t2\  +  2t\)-(i  +  6t\  +  6t2k)dt 
Jo 

=  /  (t  +  m3  +  nt5)dt  =  (\t2  +  \tA  +  2tb)\\  =  i. 

Jo 


As  with  scalar  line  integrals,  we  may  define  fF-ds  when  x  is  only  a  piece- 
wise  C1  path  by  breaking  up  the  integral  in  a  suitable  manner. 
Definition  1.2  is  important  for  the  following  reason: 


Physical  interpretation  of  vector  line  integrals.  Consider  F  to  be  a  force 
field  in  space.  Then  fx  F  •  ds  may  be  taken  to  represent  the  work  done  by  F 
on  a  particle  as  the  particle  moves  along  the  path  x. 


(straight-line 
path) 


Figure  6.3 

in  moving  a 
in  a  straight 


The  work  done  by  F 
particle  from  A  to  5 
line  is  F  •  As. 


To  justify  this  interpretation,  first  recall  that,  if  F  is  a  constant  vector  field 
and  x  is  a  straight-line  path,  then  the  work  done  by  F  in  moving  a  particle  from 
one  point  along  x  to  another  is  given  by 

Work  =  F  •  As, 

where  As  is  the  displacement  vector  from  the  initial  to  the  final  position.  (See 
Figure  6.3.) 

In  general,  the  path  x  need  not  be  straight  and  the  force  field  F  may  not 
be  constant  along  x.  Nonetheless,  along  a  short  segment  of  path,  x  is  nearly 
straight  and  F  is  roughly  constant,  by  continuity.  Partition  [a,  b]  as  usual  (i.e., 
take  a  =  to  <  ■  ■  ■  <  tk  <  ■  ■  ■  <  t„  =  b)   and  focus  on  the  &th  segment. 


412       Chapter  6  |  Line  Integrals 


F(x©) 


'Ax  j. 


x(f*-i)  x($ 


Figure  6.4  Approximating  the 
work  along  a  segment  of  the 
path  x. 


(See  Figure  6.4.)  Then 

Work  done  along  &th  segment     F(x(f^ ))  •  Ax*. 

Since 

x(r  +  AO-x(f)  Ax 
x(f)  =  lim   =  hm  — , 

Af^O  Af  A^O  Af 

we  must  have,  for  Af*  =  h  —  f^-i  ~  0  and  f*_i  <  f|  <  4,  that 

x  (*D  %    =  "T  ■ 

tk  —  h-\  Af* 

Hence, 


Total  work  %  £  F(x(fj?))  •  Ax*  «      F(x(rt*))  •  x'(%*)Af*. 


jfc=i 


Therefore,  it  makes  sense  to  take  the  limit  as  all  the  Af^'s  tend  to  zero  and  define 
the  total  work  by 

// 

Work  =    lim    V  F(x(f))  •  x'(tt)Atk 

all  Art^0£— ^ 


=  f  F(x(t))-x'(t)dt 

J  a 


F-ds. 


Other  Interpretations  and  Formulations   

Suppose  x:  [a,  b]  —>  R"  is  a  C1  path  with  x'(f)  7^  0  for  a  <  t  <  b.  Recall  from 
§3.2  that  we  define  the  unit  tangent  vector  T  to  x  by  normalizing  the  velocity: 

At) 


T(/)  = 


\At)\ 


We  may  insinuate  this  tangent  vector  into  the  vector  line  integral  as  follows:  From 
Definition  1.2,  we  have 

fv-ds=  f  F(x(t))-x'(t)dt. 

Jx  J  a 


Thus, 


f  F  •  ds  =  jf  *  F(x(0)  •  |T^|  ||x'(0||  dt 


=  /  (F(x(O)-T(O)  h'(t)\\dt 


I 

L 


(F-T)ds. 


(3) 


Since  the  dot  product  F  •  T  is  a  scalar  quantity,  we  have  written  the  original  vector 
line  integral  as  a  scalar  line  integral.  We  also  see  that  fxF-ds  represents  the 


6.1  |  Scalar  and  Vector  Line  Integrals       41  3 

{scalar)  line  integral  of  the  tangential  component  ofF  along  the  path — that  is, 
how  much  of  the  vector  field  the  underlying  curve  actually  "sees."  (See  Figure  6.5.) 

An  important  interpretation  of  the  vector  line  integral  occurs  when  x  is  a 
closed  path  (i.e.,  when  x(a)  =  x(b)).  In  this  circumstance,  the  quantity  fxF  -ds 
is  called  the  circulation  of  F  along  x.  To  understand  this  idea  better,  suppose 
that  F  represents  the  velocity  vector  field  of  a  fluid.  Consider  the  amount  of  fluid 
moved  tangentially  along  a  small  segment  of  the  path  x  during  a  brief  time  interval 
At.  (We  use  r  to  denote  time  here  so  as  not  to  conflict  with  the  use  of  t  for  the 
parameter  variable  of  the  path  x.)  Since  F  •  T  gives  the  tangential  component  of 
F,  we  have  that 

Amount  of  fluid  moved  «  (F(x(O)Ar  .  T(f))  As,  (4) 

where  As  is  the  length  of  the  segment  of  the  closed  path  x.  (See  Figure  6.6.) 
Formula  (4)  is  only  approximate  because  the  segment  of  the  path  need  not  be 


Figure  6.6  The  amount  of  fluid 
transported  tangentially  along  a  segment 
of  the  closed  path  x  is  approximately 
(F(x(f))Ar-T(?))  As. 


completely  straight  (so  that  T(t)  may  not  be  a  constant  vector),  and  also  because 
the  vector  field  F  need  not  be  constant  over  the  segment  of  the  path.  If  we  divide 
the  term  in  (4)  by  At,  then  the  average  rate  of  transport  along  the  segment 
during  the  time  interval  At  is  (F(x(/))  •  T(f))  As.  If  we  now  partition  the  closed 
path  x  into  finitely  many  such  small  segments  and  sum  the  contributions  of  the 
form  (F(x(f ))  •  T(f ))  As  for  each  segment,  then  let  all  the  lengths  As  tend  to 
zero,  we  find  that  the  average  rate  of  fluid  moved,  denoted  AL/At,  is  given 
approximately  by 


Finally,  if  we  let  At 
dL/dr,  to  be 


/  \W  t 

Figure  6.5  The  vector  line  integral 
f^F-ds  equals  /x(F  •  T) ds,  the 
scalar  line  integral  of  the  tangential 
component  of  F  along  the  path. 


AL  f 
—  «  /  (F-T)Js. 

^  Jx 


Ai 

0,  we  may  define  the  (instantaneous)  rate  of  fluid  moved, 


—  =  [  (F-F)ds  =  [  F-ds, 

at     A  Jx 


which  is  what  we  have  called  the  circulation. 


414       Chapter  6  I  Line  Integrals 


F  =  xi  +  yj 

"\*2  +  y2  = 

Figure  6.7  At  every  point  along 
the  circle,  F  =  x  i  +  y  j  has  no 
tangential  component. 


EXAMPLE  4   The  circle  x2  +  y2  =  9  may  be  parametrized  by  x  =  3  cos  t,  y  = 

3  sinf,  0  <  t  <  2n,  Hence,  a  unit  tangent  vector  is 

-3  sinf  i  +  3  cosf  j 


T  = 


sinz  t  +  9  cos2  / 


=  —  sin  t\  +  cos  t  j  = 


Now  consider  the  radial  vector  field  F  =  x  i  +  y  j  on  R2.  At  every  point  along  the 
circle  we  have 


-y\  +  x\ 


=  0. 


F-T  =  (xi  +  yj) 

Thus,  F  is  always  perpendicular  to  the  curve.  Therefore, 

f¥-ds=  j  (F-T)ds  =  j  Ods  =  0 

and  considering  F  as  a  force,  no  work  is  done.  Considering  F  as  a  velocity  field, 
the  circulation  of  F  along  x  is  likewise  zero.  (See  Figure  6.7.) 
On  the  other  hand  if  instead  F  =  y  i  —  2x  j,  then 

■•2  2x2 


F.T  =  G>i-2*j). 


-yi  +  xj 


-y 


3 


This  quantity  will  always  be  negative  on  the  circle,  so  we  expect  the  circulation 
to  be  negative.  In  particular,  we  have 

f¥-ds=  f(F-T)ds=  f --  +2X  ds. 

Jx  Jx  Jx  3 

Using  the  parametrization  given  above,  we  have  ||x'(f)||  =  3,  so  that 


L 


y 


2x2 


9sin2f  +  18  cos2/ 


•  3dt 


-L 

=  -9       (sin2  t  +  2  cos2  t)dt  =  -9       (1  +  cos2 1)  dt 
Jo  Jo 

=  -9        (1  +  |(1  +cos2?))  dt 
Jo 


=  -9(|f +  isin2/)|^  =  -27tt.  ♦ 

Next,  suppose  that  x(?)  =  (x(f),  y(t),  z(t)),  a  <  t  <  b,  is  a  C1  path  and 

F(x,  y,  z)  =  M(x,  y,  z)i  +  N(x,  y,  z)j  +  P(x,  y,  z)k 

is  a  continuous  vector  field.  Then,  from  Definition  1 .2  of  the  vector  line  integral, 
we  have 

f¥-ds=  f  (M(x,y,z)i  +  N(x,y,z)]  +  P(x,y,z)k) 

Jx  J  a 

■  (x'(t)i  +  y'(t)j  +  z'(t)K)dt 

=  f  [M(x,  y,  z)x'(t)  +  N(x,  y,  z)y'{t)  +  P(x,  y,  z)z'(t)]  dt. 

Recall  that  the  differentials  of  x,  y,  and  z  are 

dx  =  x'(t)dt,        dy  =  y'(t)dt,        dz  =  z\t)dt. 


6.1  |  Scalar  and  Vector  Line  Integrals       41  5 


Hence, 


¥-ds=  /  M(x,y,z)dx  +  N(x,y,z)dy  +  P(x,y,z)dz. 

Jx  Jx 


The  integral 

M  dx  +  N  dy  +  P  dz 


L 


is  a  notational  alternative  to  f  F  •  ds.  Indeed,  the  former  is  defined  by  the  latter. 
The  integral 


L 


M  dx  +  N  dy  +  P  dz 


is  commonly  referred  to  as  the  differential  form  of  the  line  integral.  (In  fact,  the 
expression  M  dx  +  N  dy  +  P  dz  is  itself  called  a  differential  form.)  It  empha- 
sizes the  component  functions  of  the  vector  field  F  and  arises  regularly  in  applica- 
tions. Be  sure  to  interpret  M  dx  +  N  dy  +  P  dz  carefully:  It  should  be  evaluated 
using  the  parametric  equations  for  x,  y,  and  z  that  come  from  the  path  x. 


EXAMPLE  5   We  compute 

(y  +  z)  dx  +  (x  +  z)  dy  +  (x  +  y)  dz, 

where  x  is  the  path  x(f)  =  (t,  t2,  t3)  for  0  <  t  <  1. 

Along  the  path,  we  have  x  =  t,  y  —  t2,  and  z  =  t3  so  that  dx  =  dt,  dy  = 
2t  dt,  and  dz  =  3>t2  dt.  Therefore, 


(y  +  z)  dx  +  (x  +  z)  dy  +  (x  +  y)  dz 


=  f  (t2  +  t3)  dt  +  (t  +  t3)2t  dt  +  (t  +  t2)3t2  dt 
Jo 


L 


=  /  (5t« +  4t* +  3tz)dt  =  (t:> +  t4 +  ti)\:.  =  3. 


Line  Integrals  Along  Curves: 

The  Effect  of  Reparametrization  

Since  the  unit  tangent  vector  to  a  path  depends  on  the  geometry  of  the  underlying 
curve  and  not  on  the  particular  parametrization,  we  might  expect  the  line  integral 
likewise  to  depend  only  on  the  image  curve.  We  shall  see  precisely  to  what  degree 
this  observation  is  true  generally  for  both  vector  and  scalar  line  integrals. 
We  begin  with  an  example.  Consider  the  following  two  paths  in  the  plane: 

x:  [0,  2n]      R2,    x(t)  —  (cost,  sinf) 

and 

y:[0, 7r]  ^  R2,    y(0  =  (cos2*,  sin2f). 


416       Chapter  6  I  Line  Integrals 


It  is  not  difficult  to  see  that  both  x  and  y  trace  out  a  circle  once  in  a  counterclockwise 
sense.  In  fact,  if  we  let  u(t)  =  2t,  then  we  see  that  y(r)  =  x(m(/)).  That  is,  the  path 
y  is  nothing  more  than  the  path  x  together  with  a  change  of  variables,  which 
suggests  the  following  general  definition: 


DEFINITION  1.3  Let  x:  [a,  b]  ->  R"  be  a  piecewise  C1  path.  We  say  that 
another  C1  path  y:  [c,d]  — >  R"  is  a  reparametrization  of  x  if  there  is 
a  one-one  and  onto  function  u:  [c,  d]  ->  [a,  b]  of  class  C1,  with  inverse 
u~x :  [a,  b]  [c,  d]  that  is  also  of  class  C1,  such  that  y(f)  =  x(m(?));  that  is, 
y  =  x  o  u. 


Reflecting  on  Definition  1.3  should  convince  you  that  any  reparametrization 
of  a  path  must  have  the  same  underlying  image  curve  as  the  original  path. 

EXAMPLE  6   The  path 

x(0  =  (1  +  It,  2  -  t,  3  +  50,  0<f<l, 

traces  the  line  segment  from  the  point  (1,  2,  3)  to  the  point  (3,  1,  8).  So  does  the 
path 

y(r)  =  (l  +  2?2,2-f2,3  +  5f2),    0  <  t  <  1. 

We  have  that  y  is  a  reparametrization  of  x  via  the  change  of  variable  u(t)  =  t2. 
However,  the  path  z:  [—  1 ,  1]  ->  R3  given  by 

z(0  =  (l+2f2,2-r2,3  +  5f2) 

is  not  a  reparametrization  of  x.  We  have  z(f)  =  x(u(t)),  where  u(t)  =  t2,  but  in 
this  case  u  maps  [—  1 ,  1]  onto  [0,  1]  in  a  way  that  is  not  one-one. 
On  the  other  hand, 

w(0  =  (3  -  It,  1  + 1,  8  -  50,    0  <  /  <  1, 

is  a  reparametrization  of  x.  We  have  w(0  =  x(l  —  t),  so  the  function  u:  [0,  1]  — > 
[0,  1]  given  by  u(t)  =  1  —  t  provides  the  change  of  variable  for  the  reparametriza- 
tion. Geometrically,  w  traces  the  line  segment  between  (1,  2,  3)  and  (3,  1,  8)  in 
the  opposite  direction  to  that  of  x.  ♦ 

If  y:  [c,  d]  — >  R"  is  a  reparametrization  of  x:  [a,  b]  — >  R"  via  the  change  of 
variable  u:  [c,  d]  — >•  [a,  b],  then,  since  u  is  one-one,  onto,  and  continuous,  we 
must  have  either 

(i)  u(c)  =  a  and  u(d)  =  b,  or 

(ii)  u(c)  =  b  and  u(d)  =  a. 

In  the  first  case,  we  say  that  y  (or  u)  is  orientation-preserving  and,  in  the  second 
case,  that  y  (or  u)  is  orientation-reversing.  The  idea  is  that  if  y  is  an  orientation- 
preserving  reparametrization,  then  y  traces  out  the  same  image  curve  in  the  same 
direction  that  x  does,  and  if  y  is  orientation-reversing,  it  traces  the  image  in  the 
opposite  direction. 


6.1  |  Scalar  and  Vector  Line  Integrals  417 


EXAMPLE  7  Ifx:  [a,  b]  — >  R"  is  any  C1  path,  then  we  may  define  the  opposite 
path  xopp:[a,b]  ->  R"  by 

xopP(0  =  x(a  +  b  -  t). 

(See  Figure  6.8.)  That  is,  xopp(f)  =  x(u(t)),  where  u:  [a,  b]  ->  [a,  £>]  is  given  by 
w(f)  =  a  +  b  —  t.  Clearly,  then,  xopp  is  an  orientation-reversing  reparametrization 
ofx.  ♦ 

Figure  6.8  A  path  and  its 

opposite.  jn  addition  to  reversing  orientation,  a  reparametrization  of  a  path  can  change 

the  speed.  This  follows  readily  from  the  chain  rule:  If  y  =  x  o  u,  then 

y'(t)=^(x(u(t)))  =  x'(u(t))u'(t).  (4) 

So  the  velocity  vector  of  the  reparametrization  y  is  just  a  scalar  multiple  (namely, 
w'(0)  of  the  velocity  vector  of  x.  In  particular,  we  have 

Speed  ofy=  ||y'(0||  =  ||  u'(t)  x'("(0)  || 

=  \u'(t)\  ||x'(m(0)||  =  \u'(t)\  -  (speed  ofx).  (5) 

Since  u  is  one-one,  it  follows  that  either  u'(t)  >  0  for  all  t  6  [a,  b]  or  u'(t)  <  0 
for  all  t  6  [a,  b].  The  first  case  occurs  precisely  when  y  is  orientation-preserving 
and  the  second  when  y  is  orientation-reversing. 

How  does  the  line  integral  of  a  function  or  a  vector  field  along  a  path  differ 
from  the  line  integral  (of  the  same  function  or  vector  field)  along  a  reparametriza- 
tion of  a  path?  Not  much  at  all.  The  precise  results  are  stated  in  Theorems  1 .4 
and  1.5. 


THEOREM  1.4  Let  x:  [a,b]  R"  be  a  piecewise  C1  path  and  let  f:X  C 
R"  —>  R  be  a  continuous  function  whose  domain  X  contains  the  image  of  x. 
If  y:  [c,  d]  —>  R"  is  any  reparametrization  of  x,  then 


f  fds=  [ 

Jy  Jx 


fds. 


PROOF  We  will  explicitly  prove  the  result  in  the  case  where  x  (and  therefore,  y) 
is  of  class  C1.  (When  x  is  only  piecewise  C1,  we  can  always  break  up  the  integral 
appropriately.)  In  the  C1  case,  we  have,  by  Definition  1 . 1  and  the  observations  in 
equation  (5),  that 

f  fds=  fdf(y(t))\\y(t)\\dt=  f  f(x(u(t)))\\x'{u(t))\\\u'(t)\dt. 

If  y  is  orientation-preserving,  then  u(c)  =  a,  u(d)  =  b,  and  \ur(t)\  =  uf(t).  Thus, 
using  substitution  of  variables, 

f(x(u(t)))\\x'(u(t))\\  \u\t)\dt=  J  f(x(u(t)))\\x'(u(t))\\u'(t)dt 

=  /  f(x(u))  ||x'(w)||  du  =  /  fds. 

J  a  Jx 


418       Chapter  6  |  Line  Integrals 


If,  on  the  other  hand,  y  is  orientation-reversing,  then  u(c)  =  b,  u{d)  =  a,  and 
\u'(t)\  =  —u'(t),  since  u'(t)  <  0.  Therefore,  in  this  case, 

/d  pd 
f(x(u(t)))  \\x'(u(t))\\  \u'(t)\ dt  =      f(x(u(t)))  ||x'(w(/))||  (-u'(t))dt 

=  —  I  f(x(u))\\x'(u)\\du 
Jb 

—  I  f(x(u))  ||X'(M)||  du  =  I  fds. 

J  a  Jx 

Hence,  f  f  ds  =  fxf  ds  in  either  case,  as  desired.  ■ 

THEOREM  1.5  Letx:  [a,  b]  -»  R"  beapiecewiseC1  pathandletF:  X  C  R"  -+ 
R"  be  a  continuous  vector  field  whose  domain  X  contains  the  image  of  x.  Let 
y:  [c,  d]  — >  R"  be  any  reparametrization  of  x.  Then 

1.  If  y  is  orientation-preserving,  then  f  F  •  ds  =  fx  F  •  ds. 

2.  If  y  is  orientation-reversing,  then  f  F  •  ds  =  —fx  F  •  ds. 


PROOF  As  in  the  proof  of  Theorem  1.4,  we  consider  only  the  C1  case  in  detail. 
Using  Definition  1 .2  for  the  vector  line  integral,  equation  (4),  and  substitution  of 
variables,  we  have 

n  nd  nd 

F-ds=      ¥(y(t))-y'(t)dt=  F(x(u{t)))-x\u(t))u'(t)dt. 

Jy  J c  J c 

If  y  is  orientation-preserving,  then  u(c)  =  a,  u(d)  =  b,  so  we  have 

r  d  pb  p 

/    F(\(u(t)))-x'(u(t))u'(t)dt  =      F(x(u))-x'(u)du  =  /  F-  ds. 

J c  J  a  Jx 

This  proves  part  1.  If  y  is  orientation-reversing,  then  u(c)  =  b,  u(d)  =  a,  so, 
instead,  we  have 

p  d  pa  p 

l    F(x(m(0))  •  x'(m(0)  u'(t)  dt  =      F(x(w))  •  x'(m)  du  =  -  /  F  •  ds, 

Jc  Jb  Jx 

which  establishes  part  2.  ■ 

Simply  put,  Theorems  1 .4  and  1 .5  tell  us  that  scalar  line  integrals  are  entirely 
independent  of  the  way  we  might  choose  to  reparametrize  a  path.  Vector  line 
integrals  are  independent  of  reparametrization  up  to  a  sign  that  depends  only  on 
whether  the  reparametrization  preserves  or  reverses  orientation. 

EXAM  PLE8  LetF  =  xi  +  yj,  and  consider  the  following  three  paths  between 
(0,  0)and(l,  1): 


x(0  =  (t,t)  0  <  t  <  1, 

l 

2  ' 


y(0  =  (2f,20  0  <r<i, 


and 

z(0  =  (!-*,  1-0  0<?<1. 


6.1  |  Scalar  and  Vector  Line  Integrals       41  9 


The  three  paths  are  all  reparametrizations  of  one  another;  x,  y,  and  z  all  trace  the 
line  segment  between  (0,  0)  and  (1,  1) — x  and  y  from  (0,  0)  to  (1,  1)  and  z  from 
(1,  l)to(0,  0). 

We  can  compare  the  values  of  the  line  integrals  of  F  along  these  paths: 


j^F-ds  =  F(x(t))-x'(t)dt 

=  [  (ti  +  tj)-(i  +  ])dt  =  f  2tdt=t2\lQ=U 
Jo  Jo 

[F-ds=  [  '  (2M  +  2rj)-(2i  +  2j)d/  =  f 

Jy  JO  JO 

((l-r)i  +  (l-0j)-(-i-j)^ 


1/2  rl/2  \n 

%tdt  =  4t2\0'  =  1; 


lF-ds  =  l 


■  i 

\2| 


[  2(t-Y)dt  = 
Jo 


The  results  of  these  calculations  are  just  what  Theorem  1.5  predicts,  since  y  is 
an  orientation-preserving  reparametrization  of  x  and  z  is  an  orientation-reversing 
one.  ♦ 

The  significance  of  Theorems  1.4  and  1.5  is  not  merely  that  they  allow  us 
occasionally  to  predict  the  results  of  line  integral  computations  but  also  that  they 
enable  us  to  define  line  integrals  over  curves  rather  than  over  parametrized  paths. 
To  be  more  explicit,  we  say  that  a  piecewise  C1  path  x:  [a,  b]  ->  R"  is  closed 
if  x(a)  =  x(b).  We  say  that  the  path  x  is  simple  if  it  has  no  self-intersections, 
that  is,  if  x  is  one-one  on  [a,  b],  except  possibly  that  x(a)  may  equal  x(b).  Then, 
by  a  curve  C,  we  now  mean  the  image  of  a  path  x:  [a,  b]  ->  R"  that  is  one- 
one  except  possibly  at  finitely  many  points  of  [a,  b];  the  (nearly)  one-one  path  x 
will  be  called  a  parametrization  of  C.  Additionally,  we  will  refer  to  the  curve 
C  as  being  closed  or  simple  if  it  has  a  corresponding  parametrization  that  is 
closed  or  simple.  (See  Figure  6.9.)  It  is  a  fact  (whose  proof  we  omit)  that  if  x 
and  y  are  both  parametrizations  of  the  same  simple  curve  C,  then  they  must  be 
reparametrizations  of  each  other. 


Not  simple,  not  closed 


Simple,  not  closed 


Not  simple,  closed 


Simple,  closed 


Figure  6.9  Examples  of  curves. 


420       Chapter  6  |  Line  Integrals 


(0,3) 

.J 

Figure  6.10  The  ellipse 
^2/25  +  y2/9  =  1  of  Example  9 
is  parametrized  in  the  counter- 
clockwise sense  by  x(f )  = 
(5  cost,  3  sinf),  0  <  t  <  lit,  and 
in  the  clockwise  sense  by  y(r)  = 
(5  cos2(;r  —  t),  3  sin2(7r  —  t)), 
0  <  t  <  jr. 


or 


CO 

or 

CO 


Figure  6.1 1  The  possible 
orientations  for  a  nonclosed  and 
a  closed,  simple  curve. 


EXAMPLE  9  The  ellipse  x2/25  +  y2/9  =  1  shown  in  Fi  gure  6. 1 0  is  a  simple, 
closed  curve  that  may  be  parametrized  by  either 

x:  [0,  2jt]  ->  R2,    x(f)  =  (5cosf,  3 sin?) 

or  by 

y:[0,  7t]  ->  R2,    y(f )  =  (5  cos 2(jt  -  f).  3sin2(jr  - /)), 

since  both  paths  have  the  ellipse  as  image  and  each  map  is  one-one,  except  at  the 
endpoints  of  their  respective  domain  intervals.  Note  that  y  is  a  reparametrization 
of  x. 

However,  the  path 

z:  [0,  6jt]  ->  R2,     z(r)  =  (5cosf,  3  sinf) 

is  not  a  parametrization  of  the  ellipse,  since  it  traces  the  ellipse  three  times  as  t 
increases  from  0  to  6jt.  That  is,  z  is  not  one-one  on  (0,  6jt).  ♦ 

Simple  curves,  whether  closed  or  not,  always  have  two  orientations  that  cor- 
respond to  the  two  possible  directions  of  travel  along  any  parametrizing  path.  (See 
Figure  6.1 1.)  We  say  that  a  simple  curve  C  is  oriented  if  a  choice  of  orientation 
is  made. 

The  reason  for  the  preceding  terminology  is  that  if  C  is  a  (piecewise  C1) 
simple  curve,  we  may  unambiguously  define  the  scalar  line  integral  of  a  continuous 
function  /  over  C  by 


Lfds  =  l 


fds, 


where  x  is  any  parametrization  of  C.  Theorem  1.4  guarantees  that  the  choice  of 
how  to  parametrize  C  will  not  matter. 

On  the  other  hand,  we  can  define  the  vector  line  integral  only  for  oriented 
simple  curves.  If  an  orientation  for  C  is  chosen  and  x:  [a,  b]  — >  R"  is  a  parame- 
trization of  C  that  is  consistent  with  this  orientation,  then  we  define  the  vector 
line  integral  of  a  continuous  vector  field  F  over  C  by 


F  •  ds. 


Theorem  1.5  shows  that  this  definition  is  independent  of  the  choice  of  para- 
metrization of  C,  as  long  as  it  is  made  consistently  with  the  given  orienta- 
tion. Indeed  if  C  denotes  the  same  curve  as  C,  only  oriented  in  the  opposite 
way,  then  C  is  parametrized  by  xopp  (where  x  parametrizes  C)  and  we  have, 
by  Theorem  1.5, 


I    F-ds=  [    F-ds  =  -  J 


F-Js 


-I 


F-ds. 


EXAMPLE  10  Let  C  be  the  upper  semicircle  of  radius  2,  centered  at  (0,  0) 
and  oriented  counterclockwise  from  (2,  0)  to  (—2,  0).  Then  we  may  calculate 
Jc(x2  —  y2  +  1)  ds  by  choosing  any  parametrization  for  C.  For  instance,  we  may 
parametrize  C  by 


x(t)  =  (2 cos?,  2  sinf),    0  <  t  <  it 


6.1  |  Scalar  and  Vector  Line  Integrals  421 


J  (x2-y2+l)d 


or  by 

y(t)  =  (-2cos2f,  -2sin2f),     -7t/2  <t<0. 
(Note  that  y(t)  =  x(2t  +  n).)  Then 

j  (x2  -  y2  +  l)ds  =  j{x2  -y2  +  l)ds 

=  /   (4  cos2  t  -  4  sin2  t  +  1)V 4  sin2 1  +  4  cos2 1  dt 
Jo 

=  f  (4cos2/  +  1)2 dt  =  2(2sin2?  +  t)\n  =  2it, 
Jo 

by  the  double-angle  formula.  Similarly,  we  check 

=  /     (4  cos2  2t  -  4  sin2  2r  +  1)V  16  sin2  2?  +  16  cos2  2t  dt 

J-n/2 

/o 
(4cos4f  +  1)  •  4dt  =  4(sin4f  +  0|°  /2  =  2jr- 
-jr/2  " 


(2, 12, 0)         EXAMPLE  11    We  calculate  the  work  done  by  the  force 

F  =  xi  -  yj  +  (x  +  y  +  z)k 

Figure  6.1 2  The  oriented  curve  on  a  particle  that  moves  along  the  parabola  y  =  3x2,  z  =  0  from  the  origin  to  the 
of  Example  11.  point  (2,  12,  0).  (See  Figure  6.12.) 

We  parametrize  the  parabola  by  x  =  t,  y  =  3t2,  z  =  0  for  0  <  t  <  2.  Then, 
by  Definition  1 .2, 

Work: 

IC  Jx  JO 


=  f  F-ds=  f¥-ds=  f  F(x(t))-x'(t)dt 

Jc  Jx  Jo 

=  f  (t,-3t2,t  +  3t2)-(l,6t,0)dt  =  f  (t-m^)dt 
Jo  Jo 


=  {\t2-\t%  =  2-12=-lQ. 

The  meaning  of  the  negative  sign  is  that  by  moving  along  the  curve  in  the  indicated 
direction,  work  is  done  against  the  force.  If  we  orient  the  curve  the  opposite 
way,  however,  then  the  work  done  in  moving  from  (2,  12,  0)  to  (0,  0,  0)  would 
be  70.  ♦ 


Numerical  Evaluation  of  Line  Integrals  (optional)   

If  we  have  a  scalar  line  integral  fc  f  ds,  or  a  vector  line  integral  fc¥-ds,  and  a 
suitable  parametrization  x  of  C,  then  Definitions  1 . 1  and  1 .2  enable  the  evaluation 
of  the  line  integrals  by  means  of  definite  integrals  in  the  parameter  variable.  These 
definite  integrals  may  be  difficult — or  even  impossible — to  evaluate  exactly,  so 


422       Chapter  6  |  Line  Integrals 


we  might  need  to  resort  to  a  numerical  method  (such  as  the  trapezoidal  rule  or 
Simpson's  rule)  to  approximate  the  value  of  the  integral. 


Ao  1 

Figure  6.1  3  The  curve  C  with  n 
points  chosen  on  it. 


If  a  (vector)  line  integral  is  given  in  differential  form  as  fc  M  dx  +  N  dy  + 
P  dz  (or  as  fc  M  dx  +  N  dy  if  we  are  working  in  R2),  then  we  can  give  numerical 
approximations  without  resorting  to  any  parametrization  as  follows.  First,  choose 
points  xo,  Xi , . . .  x„  along  C,  where  xo  is  the  initial  point  and  x„  is  the  terminal 
point.  (See  Figure  6.13.)  For  k  —  0, . . . ,  n,  let  us  write 

x*  =  (xk,  yic,  Zk) 

and  for  k  =  1, . . . ,  n,  let 

Axk  =xk-  xk-i,  Ayk  =  yk  -  yt-u  Azk  =  Zk  -  Zfc-i- 

Finally,  let  xk  =  (xk,  yk,  z*k)  denote  any  point  on  the  arc  of  C  between  x^-i  and 
%k-  Then  we  may  approximate  the  line  integral  as 

/  M  dx  +  N  dy  +  P  dz  «  V  [M(x%)Axk  +  N(x*k)Ayk  +  P(x*k)Azk] .  (6) 
Jc  k=i 

Besides  the  approximation  given  in  (6),  we  may  also  use  a  version  of  the 
trapezoidal  rule  adapted  to  the  case  of  line  integrals.  In  particular,  the  trapezoidal 
rule  approximation  Tn  to  the  line  integral  jc  M  dx  +  N  dy  +  P  dz  is 


k=l  L 


(M(x*_o  +  M(x,))  ^  +  (^(^-0  +  ^(x*)) 


+  (P(x,_0+JP(x,))^ 


fc=l  k=\ 


=      (M(xk-0  +  M(xk))  ^  +      (Mxk-i)  +  N(xk))  Al 


n  a 

+  J](JP(x,_i)  +  P(x,))-^.  (7) 
k=\  1 


6.1  |  Scalar  and  Vector  Line  Integrals  423 


Similar  (and  shorter)  formulas,  analogous  to  (6)  and  (7),  may  be  given  to  approx- 
imate the  two-dimensional  line  integral  fc  M  dx  +  N  dy;we  will  not  state  them 
explicitly. 

EXAMPLE  12  Let  C  be  the  ellipse  x2  +  4y2  =  4,  oriented  counterclockwise. 
(See  Figure  6.14.)  We  approximate  fc  y2  dx  +  x  dy  using  the  (two-dimensional 
version  of)  the  trapezoidal  rule  (7). 


,x1 

yo  =  x4 

x3 

Figure  6.14  The  ellipse  C  in 
Example  12. 

To  make  this  approximation,  let 
x0  =  (2,  0),  xi  =  (0,  1),  x2  =  (-2,  0),  x3  =  (0,-1),  x4  =  (2,  0). 
Then  we  have 

Axi  =  A^2  =  —2,  AX3  =  AX4  =  2; 

A^  =  1,  Ay2  =  Ay3  =  -1,  Ay4  =  1. 
Hence,  from  (7),  we  compute 

TA  =  (02  +  l2)f  +  (l2  +  02)^  +  (02  +  (-l)2)f  +  ((-l)2  +  02)§ 

+  (2  +  0)i  +  (0  +  (-2))^  +  ((-2)  +  0)=±  +  (0  +  2)\ 
=  -1-1  +  1  +  1  +  1  +  1  +  1  +  1=4. 
The  exact  answer  may  be  found  using  the  parametrization 

Ix  =  2cosf 
y  =  sint 

We  calculate  that 


0  <  t  <  2tt. 


I  y2  dx  +  x  dy  =  /  (sin2  t(—  2  sinf)  +  2  cos2  f)  dt 
Jc  Jo 


-2tj 


f 

/  (2(1  -  cos2  f)(- sinf)  +  1  +  cos2f)  dt 
Jo 


=  (2 


cos  t 


\  cos3 1  +  t  +  \  sin2r)  |^  =  2jt. 


From  this  we  see  that  our  approximation  is  quite  rough.  It  can  be  improved 
by  increasing  n  while  taking  smaller  values  of  Axk  and  Ay^.  For  instance  (see 


424       Chapter  6  |  Line  Integrals 


Figure  6.15  The  ellipse  C  in  Example 
12  with  eight  points  marked. 


Figure  6.15),  if  we  let 
x0  =  (2,  0),  Xl  =  (V3,  i)  ,  x2  =  (0,  1),  x3  =  (-73,  |)  ,  X4  =  (-2,  0), 

x5  =  (-V3,  -i)  ,  x6  =  (0,  -1),  x7  =  (y/3,  -i)  ,  x8  =  (2,  0), 


then 


Axi  =  V3  -  2,  A^2  =  Ajf3  =  -V?>,  AX4  =  ^3  -  2,  Ax5  =  2  -  V^, 
Ax6  =  Ax7  =  V^,  Ax8  =  2  -  V^; 
Ay!  =  Ay2  =  L  Ay3  =  Ay4  =  Ay5  =  Ay6  =  -\,  Ay7  =  Ay8  =  \. 


Therefore,  fc  y2  dx  +  x  dy  is  also  approximated  by 


T*=(02  +  (tf)^  +  ((tf+l2)^  +  ((l2+2-)2) 


V3 


-V3 


2  '  "  /    2    +  Vv" 


+  (G)2  +  o2)^  +  (o2  +  (-i)2) 


+  ((4)2  +  (-D2)f  +  ((-d2  +  (4)2) 

+  ((- \f  +  02)  ^  +  (2  +  ^  +  (V3  +  0)f 
+  (0  +  (-V3))^  +  ((-  V3)  +  (-2))^ 
+  ((-2)  +  (-  V3))4^  +  (("  V3)  +  0)4^ 


2-  V3 


V3 


+  (0  +  V3)^  +  (V3  +  2)^ 

=  2  +  2^3  R»  5.4641. 
Although  still  rough,  this  represents  a  better  approximation. 


6.1  |  Scalar  and  Vector  Line  Integrals  425 


EXAMPLE  13  Let  C  be  the  line  segment  from  (0,  0,  0)  to  (2,  2,  2),  and  con- 
sider the  line  integral  fc  y  V 1  +  x 3  dx  +  e~xz  dy  +  cos  (z2)  dz.  The  standard 
parametrization  of  the  line  segment  C  by 


x  =  t 

y=t  0<t<2 

z  =  t 


enables  us,  in  theory,  to  evaluate  the  line  integral  above  by  evaluating  the 
one-variable  integral 


1  +  /3  +  e~'~  +  cos 


(t2)) 


dt. 


Unfortunately,  we  cannot  do  so  exactly  (i.e.,  by  means  of  the  fundamental 
theorem  of  calculus,  since  an  antiderivative  of  the  integrand  cannot  be  found 
in  terms  of  elementary  functions).  Instead  we  provide  approximations  using 
formulas  (6)  and  (7). 

Let  n  =  4.  If  we  choose  points  along  C  that  are  regularly  spaced  then  we 
have 

x0  =  (0,  0,  0),  Xl  =  (i,  \,\),  x2  =  (1,1,1),  *3  =  (§,§,§),  *4  =  (2>  2'  2), 
so  that 

Axk  =  Ayk  =  Azk  =  \. 
To  calculate  an  approximation  using  formula  (6),  we  can,  for  example,  let 

X^  =  l(Xjt_l  +  xk), 
which  is  the  midpoint  of  the  line  segment  joining  xj_i  and  X&.  Then 

y*  —  ( —    I    I)     x*  -  (I    1    1\     Y*  —  (—    5    5\        *  _  (I    7  7\ 
Al  —  U'  4'  4/  '     2  —  U'  4'  4/  '      3  —  U'  4'  4/  '     4  —  U'  4'  4/  ' 

and  so  formula  (6)  tells  us  that 

J  y  v7!  +  *3  dx  +  e~-vz  <iy  +  cos (z2) <iz 


+ 
+ 

+ 


i  +  G)V«-1/16§  + 


cos 


J_  1 

16  '  2 


cos 


_9_  1 
16  '  2 


+  +e-25/16  1  +CQS  25 


16  2 


yi  +  dn  +  e-^i  +  cos 


49  1 
16  '  2 


Rs  5.16422. 


426       Chapter  6  |  Line  Integrals 

Using  the  trapezoidal  rule  formula  (7),  we  may  also  approximate  the  line  inte- 
gral by 

t4  =  (o  +  kyf^W)  f  +  (lv/r+W  +  f 

+  (l/TTP  +  1^1+ (|)3)  f  +  (l7l  +  (l)3  +  2^1+2*)  f 
+  (e°  +  e->/4)^  +  (e-i/4  +  ,-1)1/?  +  (e-i  +  ,-9/4)1/2 
+  (e-9/4  +  ^-4)^  +  (cos  0  +  cos  i)^  +  (cos  ±  +  cos  1)^ 
+  (cos  1  +  cos        +  (cos  5  +  cos 4)^ 
%  5.44874. 

As  was  the  case  in  Example  12,  because  of  the  small  number  of  sampling 
points  used,  formulas  (6)  and  (7)  provide  relatively  crude  approximations.  ♦ 


6.1  Exercises 


1.  Let  fix,  y)  =  x  +  2y.  Evaluate  the  scalar  line  integral 
/  /  ds  over  the  given  path  x. 

(a)  x(?)  =  (2  -  3t,  4t  -  1),  0  <  t  <  2 

(b)  x(f)  =  (cosf,  sinf),  0  <  t  <  tt 

In  Exercises  2-7,  calculate  J  f  ds,  where  f  and  x  are  as 
indicated. 

2.  f(x,  y,  z)  =  xyz,  x(f)  =  (t,  It,  3t),  0  <  t  <  2 
x  +  z 


3.  fix,  y,  z) 


y  +  z 


,  x(t)  =  (t,  t,  f3/2),  1  <  t  <  3 


4.  fix,  y,  z)  =  3x  +  xy  +  z3,  x(f)  =  (cos4f,  sin4f,  3t), 
0  <  t  <  2jt 


5.  f(x,y,z) 


x2  +  y 


■ ,  x(f )  =  ie2t  cos  3t ,  e1'  sin  3t ,  e2'), 


0  <  t  <  5 

6.  fix,  y,z)  =  x  +  y  +  z, 

1(2*,  0,  0)  if0<f<l 
(2,3f-3,0)  if  1  <  f  <  2 
(2,  3, 2t  -4)  if2<f<3 

7.  /(A-,y,z)  =  2.x-y1/2  +  2z2, 
f(f,f2,0)        ifO  <t  <  1 


x(0 


Exercises  8— 16,  find  f  F  •  ds,  where  the  vector  field  F  and 
the  path  x  are  given. 

B.F  =  xi  +  y)  +  zK   x(t)  =  (2t  +  \,t,3t-l),  0< 
t  <  1 

9.  F  =  iy  +  2)i  +  x  j,  x(f)  =  (sinf,  -cosf),  0<f< 

JT/2 

10.  F  =  xi  +  yj,  x(f)  =  (2f  +  1,  f  +  2),  0  <  f  <  1 

11.  F  =  iy  -  jc)i  +  xVj,x(f)  =  it2,  f3),  -1  <  f  <  1 

12.  F  =  xi  +  xy]  +xyzk,     x(f)  =  (3  cos  t,  2  sinf ,  5f ), 
0  <  f  <  2tt 

13.  F  =  -3yi  +  Jcj  +  3z2k,    x(f)  =  (2f  +  1,  f2  +  f,  e'), 
0  <  f  <  1 

14.  F  =  jti  +  yj-zk,  x(f )  =  (f ,  3f2,  2f3), -1  <f  <  1 

15.  F  =  3zi  +  y2  j  +  6zk,  x(f)  =  (cos  f,  sinf ,  f/3),  0< 
f  <  4jt 

16.  F  =  y  cosz  i  +  x  sinz  j  +  jcy  sinz2  k,  x(f)  =  (f ,  f2, 
?3),  0  <  f  <  1 

1 7.  Determine  the  value  of  fxx  dy  —  y  dx,  where  x(f )  = 
(cos3f,  sin3f),  0  <  f  <  n. 


(1,1,  t-l)  ifl<f<3 


6.1  |  Exercises  427 


18.  Find  the  work  done  by  the  force  field  F  =  2xi  +  j 
when  a  particle  moves  along  the  pathx(f)  =  (t,  3t2,  2), 
0  <  t  <  2. 

19.  If    x  =  (e2fcos3f,e2'sin3/),     0  <  t  <  2n,  find 

f  x  dx  +  y  dy 

L(x2  +  y2)3/2' 

20.  Let  C  be  the  portion  of  the  curve  y  =  2^/x  between 
(1,2)  and  (9,  6).  Find  fc  3y  ds. 

21 .  Let  F  =  (x2  +  y)i  +  (y  —  x)\  and  consider  the  two 
paths 

x(f)  =  (t,  t2),  0  <  t  <  1  and 

y(f)  =  (1  -2f,4?2  -At  +  1),  0  <  t  <  5. 

(a)  Calculate  fxF-ds  and /  F  •  ds. 

(b)  By  considering  the  image  curves  of  the  paths  x  and 
y,  discuss  your  answers  in  part  (a). 

22.  Find  the  work  done  by  the  force  field  F  =  x2 y  i  +  z  j  + 
(2x  —  y)  k  on  a  particle  as  the  particle  moves  along  a 
straight  line  from  (1,  1,  1)  to  (2,  —3,  3). 

23.  Let  F  =  (2z5  -  3xy)  i  -  x2  j  +  .y2z  k.  Calculate  the 
line  integral  of  F  around  the  perimeter  of  the 
square  with  vertices  (1,  1 ,  3),  (—  1 ,  1,3),  (—  1 ,  — 1,3), 
(1,-1,3),  oriented  counterclockwise  about  the  z-axis. 

24.  Evaluate  fc(x2  —  y)  dx  +  (x  —  y2)  dy,  where  C  is  the 
line  segment  from  (1,  1)  to  (3,  5). 

25.  Find  fc  x2y  dx  —  (x  +  y)  dy,  where  C  is  the  trapezoid 
with  vertices  (0,  0),  (3,  0),  (3,  1),  and  (1,  1),  oriented 
counterclockwise. 

26.  Evaluate  fc  x2y  dx  —  xy  dy,  where  C  is  the  curve  with 
equation  y2  =  x3,  from  (1,  —1)  to  (1,  1). 

27.  Evaluate  fc  y  dx  —  x  dy,  where  C  is  the  portion  of  the 
parabola  y  =  x2,  from  (3,  9)  to  (0,  0). 

28.  Evaluate  fc(x  —  y)2  dx  +  (x  +  y)2  dy,  where  C  is  the 
portion  of  y  =  \x\,  from  (—2,  2)  to  (1,  1). 

29.  Evaluate  fc  xy2  dx  —  xy  dy,  where  C  is  the  semicir- 
cular arc  from  (0,  2)  to  (0,  —2)  traveled  clockwise. 

30.  Find  the  circulation  of  F  =  (x2  —  y)i  +  (xy  +  x)j 
along  the  circle  x2  +  y2  =  16,  oriented  counterclock- 
wise. 

31.  Evaluate  fc  yzdx  —  xzdy  +  xy  dz,  where  C  is  the 
line  segment  from  (1,  1,  2)  to  (5,  3,  1). 

32.  Calculate  fc  zdx  +  x  dy  +  y  dz,  where  C  is  the  curve 
obtained  by  intersecting  the  surfaces  z  =  x2  and 
x2  +  y2  =  4  and  oriented  counterclockwise  around  the 
z-axis  (as  seen  from  the  positive  z-axis). 


33.  Show  that  J  T  •  ds  equals  the  length  of  the  path  x, 
where  T  denotes  the  unit  tangent  vector  of  the  path. 

34.  Tom  Sawyer  is  whitewashing  a  picket  fence.  The 
bases  of  the  fenceposts  are  arranged  in  the  jcy-plane 
as  the  quarter  circle  x2  +  y2  =  25,  x,  y  >  0,  and  the 
height  of  the  fencepost  at  point  {x,y)  is  given  by 
h(x,  y)  =  10  —  x  —  y  (units  are  feet).  Use  a  scalar 
line  integral  to  find  the  area  of  one  side  of  the  fence. 
(See  Figure  6.16.) 


z 


y 


x2  +  y2  =  25 


Figure  6.16  The  picket  fence  of 
Exercise  34.  The  base  of  the  fence  is  the 
quarter  circle  x2  +  y2  =  25,  x,  y  >  0. 

35.  Sisyphus  is  pushing  a  boulder  up  a  100-ft-tall  spiral 
staircase  surrounding  a  cylindrical  castle  tower.  (See 
Figure  6.17.) 

(a)  Suppose  Sisyphus's  path  is  described  parametri- 
cally  as 

x(0  =  (5  cos  3t,  5  sin  3?,  10?),    0  <  t  <  10. 


Figure  6.17  Sisyphus's  path  up 
the  spiral  staircase  of  Exercise  35. 

If  he  exerts  a  force  with  a  constant  magnitude  of 
50  lb  tangent  to  his  path,  find  the  work  Sisyphus 


428       Chapter  6  |  Line  Integrals 


does  in  pushing  the  boulder  up  to  the  top  of  the 
tower. 

(b)  Just  as  Sisyphus  reaches  the  top  of  the  tower,  he 
sneezes  and  the  boulder  slides  all  the  way  to  the 
bottom.  If  the  boulder  weighs  75  lb,  how  much 
work  is  done  by  gravity  when  the  boulder  reaches 
the  bottom? 

36.  A  ship  is  pulled  through  a  14-ft-long  straight  channel 
by  a  line  that  passes  from  the  ship  around  a  pulley. 
Assume  that  a  coordinate  system  is  set  up  so  that  the 
pulley  is  at  the  point  (24,  32),  and  the  ship  is  pulled 
along  the  y-axis  from  the  origin  to  the  point  (0,  14).  If 
the  tension  on  the  line  is  kept  at  a  constant  25  lb,  find 
the  work  done  in  tugging  the  ship  through  the  channel. 

37.  Suppose  C  is  the  curve  y  =  f(x),  oriented  from 
(a,  f(a))  to  (b,  f(b))  where  a  <  b  and  where  /  is 
positive  and  continuous  on  [a,  b].  If  F  =  y  i,  show  that 
the  value  of  Jc  F  •  ds  is  the  area  under  the  graph  of  / 
between  x  =  a  and  x  =  b. 

38.  Let  F  be  the  radial  vector  field  F  =  xi+yj  +  zk. 
Show  that  if  x(f),  a  <  t  <  b,  is  any  path  that  lies 
on  the  sphere  x2  +  y2  +  z2  =  c2,  then  f  F  •  ds  =  0. 
(Hint:Ifx(r)  =  (x(t),  y(t),  z(t))  lies  on  the  sphere,  then 
[x(t)]2  +  [y(f)]2  +  [z(t)f  =  c2.  Differentiate  this  last 
equation  with  respect  to  t.) 

39.  Let  C  be  a  level  set  of  the  function  f(x,  y).  Show  that 
/CV/. </■  =  <>. 

40.  You  are  traveling  through  Cleveland,  famous  for  its 
lake-effect  snow  in  winter  that  makes  driving  quite 
treacherous.  Suppose  that  you  are  currently  located 
20  miles  due  east  of  Cleveland  and  are  attempting 
to  drive  to  a  point  20  miles  due  west  of  Cleveland. 
Further  suppose  that  if  you  are  s  miles  from  the  center 
of  Cleveland,  where  the  weather  is  the  worst,  you  can 
drive  at  a  rate  of  at  most  v(s)  =  2s  +  20  miles  per 
hour. 

(a)  How  long  will  the  trip  take  if  you  drive  on  a 
straight-line  path  directly  through  Cleveland?  (As- 
sume that  you  always  drive  at  the  maximum  speed 
possible.) 

(b)  How  long  will  the  trip  take  if  you  avoid  the  middle 
of  the  city  by  driving  along  a  semicircular  path 
with  Cleveland  at  the  center?  (Again,  assume  that 
you  drive  at  the  maximum  speed  possible.) 

(c)  Repeat  parts  (a)  and  (b),  this  time  using  v(s)  = 
(s2/16)  +  25  miles  per  hour  as  the  maximum 
speed  that  you  can  drive. 

41 .  Consider  a  particle  of  mass  m  that  carries  a  charge  q . 
Suppose  that  the  particle  is  under  the  influence  of  both 
an  electric  field  E  and  a  magnetic  field  B  so  that  the 
particle's  trajectory  is  described  by  the  path  x(r)  for 


a  <  t  <  b.  Then  the  total  force  acting  on  the  particle 
is  given  in  mks  units  by  the  Lorentz  force 

F  =  q(E  +  v  X  B), 

where  v  denotes  the  velocity  of  the  trajectory. 

(a)  Use  Newton's  second  law  of  motion  (i.e.,  F  =  ma, 
where  a  denotes  the  acceleration  of  the  particle) 
to  show  that 

ma  •  v  =  qTZ  •  v. 

(b)  If  the  particle  moves  with  constant  speed,  use  part 
(a)  to  show  that  E  does  no  work  along  the  path 
of  the  particle.  (Hint:  Apply  Proposition  1.7  of 
Chapter  3  to  v.) 

42.  Let  C  be  the  segment  of  the  parabola  y  =  x2  be- 
tween (0,  0)  and  (1,1)  and  consider  the  line  integral 

fc  y3  dx  —  x2  dy. 

(a)  Let  x0  =  (0,  0),  X]  =  (±,  £),  X2  =  (I ,  I),  X3  = 
(|,  yg),  X4  =  (1,  1).  Use  the  trapezoidal  rule  for- 
mula (7)  and  these  points  to  approximate  the  line 
integral. 

(b)  Now  calculate  the  exact  value  of  the  line  integral 
and  compare  your  results. 

43.  Let  C  be  the  line  segment  from  (0,  0,  0)  to  (1,  2,  3) 
and  consider  the  line  integral  fc  yz  dx  +  (x  +  z)dy  + 
x2ydz- 

(a)  Divide  the  segment  into  five  regularly  spaced 
points  xo,  xi , . . . ,  X4  and  use  the  trapezoidal  rule 
formula  (7)  to  approximate  this  line  integral. 

(b)  Compare  your  approximation  with  the  exact  value 
of  the  line  integral. 

44.  Suppose  that  magnetic  field  measurements  are  made 
along  a  wire  shaped  as  a  curve  C  and  the  results  are 
tabulated  as  follows: 


k 

Point  xt  = 

B(jct,  yk,  Zk)  = 

{Xk,  yt,  Zk) 

Mi+ N j  +  Pk 

0 

(-1,-2,-1) 

k 

1 

(0,1,-1) 

j  +  2k 

2 

(0,  2,  0) 

i  +  j  +  2k 

3 

(1.2,  1) 

2i  +  j  +  2k 

4 

(1,2,2) 

2i  +  2j  +  2k 

5 

(1,1,2) 

2i  +  3j  +  3k 

6 

(1,1,1) 

3i  +  3j  +  3k 

7 

(1,0,  0) 

4i  +  3j  +  3k 

8 

(0,  0,  0) 

4i  +  3j  +  4k 

By  writing  /  B  •  ds  as  fc  M  dx  +  N  dy  +  P  dz,  esti- 
mate the  work  done  by  B  along  C  using 

(a)  a  trapezoidal  rule  approximation; 

(b)  a  trapezoidal  rule  approximation  using  only  the 
points  x0,  x2,  x4,  x6,  x8. 


6.2  I  Green's  Theorem  429 


6.2   Green's  Theorem 

Green's  theorem  relates  the  vector  line  integral  around  a  closed  curve  C  in  R2 
to  an  appropriate  double  integral  over  the  plane  region  D  bounded  by  C.  The 
fact  that  there  is  such  an  elegant  connection  between  one-  and  two-dimensional 
integrals  is  at  once  surprising,  satisfying,  and  powerful.  Green's  theorem,  stated 
generally,  is  as  follows: 


C=  SD 


Figure  6.18  The  shaded  region  D 
has  a  boundary  consisting  of  two 
simple,  closed  curves,  each  of  class 
C1,  whose  union  we  call  C. 


THEOREM  2.1  (Green's  theorem)  Let  D  be  a  closed  bounded  region  in  R2 
whose  boundary  C  =  3D  consists  of  finitely  many  simple,  closed  piecewise  C1 
curves.  Orient  the  curves  of  C  so  that  D  is  on  the  left  as  one  traverses  C.  (See 
Figure  6.18.)  Let  F(jc,  y)  =  M(x,  y)i  +  N(x,  y)\  be  a  vector  field  of  class  C1 
throughout  D.  Then 


j>  Mdx  +  Ndy  =  ff  ( 


dN 
dx 


dM 


)dx  dy. 


(The  symbol  §c  indicates  that  the  line  integral  is  taken  over  one  or  more  closed 
curves.) 


(1,1) 


Figure  6.1 9  The  region 
D  of  Example  1. 


EXAMPLE  1  Let  F  =  xy\  +  y2]  and  let  D  be  the  first  quadrant  region  bounded 
by  the  line  y  =  x  and  the  parabola  y  =  x2.  We  verify  Green's  theorem  in  this  case. 

The  region  D  and  its  boundary  are  shown  in  Figure  6.19.  3D  is  oriented 
counterclockwise,  the  orientation  stipulated  by  the  statement  of  Green's  theorem. 
To  calculate 

®   F-ds=(f)   xydx  +  y2dy, 

JdD  JdD 

we  need  to  parametrize  the  two  C1  pieces  of  3D  separately: 


\x  =  t  x  =  1  -  t 

Ci  :  \        .     0<(<1    and    C2:  I  0  <t  <  1. 


y  =  r 


y  =  l-t 


(Note  the  orientations  of  C\  and  C2.)  Hence, 

<b  xy  dx  +  y2dy  =  J  xy  dx  +  y2dy  +  /  xy  dx  +  y2dy 
JdD  JCi  Jc2 

=  f  (t  -t2  +  t4  -2t)  dt+  f 
Jo  Jo 

=  f  (t3  +  2t5)dt+  f  2(1  -tf(-dt) 
Jo  Jo 

=  (^4+^6)i;+(!a-o3)i: 


((1  -  tf  +  (1  -  tfX-dt) 


_  1  ,  2  _  2 

~~  4  ~r  6  3 


J_ 
12' 


430       Chapter  6  |  Line  Integrals 


On  the  other  hand, 


9 

dx 


9 


dx  dy  =  I    I    —x  dy  dx  =  I    —x(x  —  x2)  dx 

J0    Jx2  JO 


=  [  (x3  -x2)dx  =  ( 
Jo 


U4_  lx3\|l 


I  _  I 

4  3 


J_ 
12  • 


The  line  integral  and  the  double  integral  agree,  just  as  Green's  theorem  says  they 
must.  ♦ 


Figure  6.20  The  disk  of 
radius  a  with  boundary 
oriented  so  that  Green's 
theorem  applies. 


Figure  6.21  The  region 
inside  the  ellipse 

x2/a2  +  y2/b2  =  1. 


EXAMPLE  2  Consider^.  —  y  dx  +  x  dy,  where  C  is  the  circle  of  radius  a  (i.e., 
the  boundary  of  the  disk  D  of  radius  a),  oriented  counterclockwise  as  shown  in 
Figure  6.20.  Although  we  can  readily  parametrize  C  and  thus  evaluate  the  line 
integral,  let  us  employ  Green's  theorem  instead: 


<j>  —y  dx  +  x  dy  =  j j 


9  9 

ir(*)  -  ^-(-y) 

ox  ay 


dx  dy 


-IL 


2  dx  dy  =  2(area  of  D)  =  lit  a 


The  rightmost  expression  is  twice  the  area  of  a  disk  of  radius  a.  In  this  case,  the 
double  integral  is  much  easier  to  consider  than  the  line  integral.  ♦ 

The  use  of  Green's  theorem  in  Example  2  can  be  put  in  a  much  more  general 
setting:  Indeed,  if  D  is  any  region  to  which  Green's  theorem  can  be  applied,  then, 
orienting  3D  appropriately,  we  have 


-y  dx  +  x  dy 


3D 


■»//. 


2  dx  dy  =  area  of  D. 


(1) 


Thus,  we  can  calculate  the  area  of  a  region  (a  two-dimensional  notion)  by  using 
line  integrals  (a  one-dimensional  construction)! 

EXAMPLE  3  Using  formula  (1),  we  compute  the  area  inside  the  ellipse 
x2/a2  +  y2/b2  =  1  (Figure  6.21). 

The  ellipse  itself  may  be  parametrized  counterclockwise  by 


\x  =  a  cos  t 


0  <  t  <  In. 


\y  =  b  sin  t 

Once  again,  using  formula  (1),  we  find  that  the  area  inside  the  ellipse  is 


i  ®    —  ydx+xdy=j  I     —bsint(—asmtdt)  +  acost(bcostdt) 
JdD  Jo 

-in 


=  ~  /     (ab  sin2 1  +  ab  cos2 1)  dt 
Jo 


ab  dt  =  Ttab. 


6.2  I  Green's  Theorem  431 


Alternative  Formulations   

We  rewrite  the  line  integral-double  integral  formula  appearing  in  the  statement 
of  Theorem  2.1  (Green's  theorem)  in  two  ways.  These  reformulations  generalize 
to  higher  dimensions  and  provide  some  additional  insight  in  interpreting  the 
geometric  significance  of  Green's  theorem. 
To  begin,  consider  a  C1  vector  field 

F  =  M(x,y)i  +  N(x,y)j 

to  be  defined  on  R3  by  taking  its  k-component  to  be  zero.  Then,  if  we  compute 
the  curl  of  F,  we  find 

\dx~~~dy~) 
Therefore,  because  k  •  k  =  1 ,  we  obtain 

Since  §SD  F  •  ds  =  §dD  Mdx  +  Ndy,  Green's  theorem  may  be  rewritten  as 
follows: 


PROPOSITION  2.2  (A  vector  reformulation  of  Green's  theorem)  If  D  is 

a  region  to  which  Green's  theorem  applies  and 

F  =  M(x,y)i  +  N(x,y)} 
is  a  vector  field  of  class  C1  on  D,  then,  orienting  3D  appropriately, 

*   F-ds  =  f  f  (Vx  F)-kdA. 

JBD  J  J D 


To  understand  this  result,  visualize  the  plane  region  D  as  sitting  in  the  xy- 
plane  in  R3.  (See  Figure  6.22.)  The  vector  k  is  a  unit  vector  normal  to  D,  and 
f /D(V  x  F)  •  kdA  is  the  double  integral  of  the  component  of  the  curl  of  F  normal 
to  D.  Since  the  line  integral  §dDF  •  ds  is  the  circulation  of  F  along  3D  (see 
§6. 1),  the  equation  of  Proposition  2.2  tells  us  that,  under  suitable  hypotheses,  the 
circulation  of  a  vector  field  F  along  the  boundary  of  a  plane  region  is  equal  to 
the  total  (or  net)  "infinitesimal  rotation"  of  F  over  the  entire  region.  (See  also 
§3.4  where  the  curl  of  a  vector  field  is  given  an  intuitive  interpretation  in  terms  of 
rotation — or  wait  until  §7.3  when  the  notion  of  curl  measuring  rotation  of  a  vector 
field  is  explained  more  precisely.)  This  result  generalizes  to  Stokes's  theorem 
in  R3.  Stokes's  theorem  relates  the  integral  of  the  component  of  the  curl  of  a 
three-dimensional  vector  field  F  that  is  normal  to  a  surface  in  R3  to  the  line 
integral  of  F  over  the  boundary  curves  of  the  surface. 

Next,  we  reformulate  Green's  theorem  in  another  way  Once  again,  assume 
that  D  is  a  region  in  R2  to  which  Green's  theorem  applies  and  that  its  boundary 
curves  are  oriented  appropriately.  At  each  point  along  the  C1  segments  of  3D, 
let  n  denote  the  unit  vector  that  is  perpendicular  to  3D  and  points  away  from  the 


V  x  F  = 


l 

d/dx 

M 


J 

d/dy 
N 


k 

3/3; 
0 


432       Chapter  6  |  Line  Integrals 


Figure  6.23  The  outward  unit 
normal  n  to  the  region  D. 


region  D.  (We  call  n  the  outward  unit  normal  vector  to  D.  See  Figure  6.23.) 
Then  we  can  demonstrate  the  following: 

THEOREM  2.3  (Divergence  theorem  in  the  plane)  If  D  is  a  region  to  which 
Green's  theorem  applies,  n  is  the  outward  unit  normal  vector  to  D,  and 

F  =  M(x,  y)\  +  N(x,  y)j 

is  a  C1  vector  field  on  D,  then 


(b    F-nds  =11  V-FdA. 

JdD  J  J D 


PROOF  Ifx(r)  =  (x(t),  y(t)),a  <  t  <  b,  parametrizes  a  C1  segment  of  3D,  then 
along  this  segment  the  unit  vector  n  may  be  obtained  geometrically  by  rotating 
the  unit  tangent  vector  x'(r)/ 1|  x'(t)  ||  clockwise  by  90°.  In  particular,  along  such  a 
parametrized  C1  segment, 

n=   /(Qi-x'(f)j  =y'(t)i-x'(t)j 

Vx'(f)2  +  fit)2  uncoil 

We  calculate  the  line  integral  <faD  F-nds.  Along  each  C1  segment  of  3D, 
the  contribution  to  the  line  integral  may  be  evaluated  as 


b 

(F(x(0)  •  n(r))  IkCOl  dt 


=  /    (M(x(t),  y(t))i+N(x(t),  y(O)j)  ■       ,  v.„     J  ||x'(0||  dt 

J  a  x  (') 


=  f  (M(x(t),  y(t))y'(t)  -  N(x(t),  y(t))x'(t))  dt 
=  I  -Ndx  +  Mdy. 

J  X 

Thus,  by  Green's  theorem, 

&  F-nds=i  -Ndx  +  Mdy  =11  [  —  -  d{-~N^  )  dA 
JdD  JdD  J  Jd\3x         dy  J 


=  11 
-11 


dM  dN\ 
dx       dy  J 


\dA 


V-FdA, 

D 

by  the  definition  of  the  divergence  of  F.  ■ 

If  C  is  a  simple,  oriented  curve,  the  line  integral  fc  F  •  nds,  where  n  is  the 
unit  normal  to  C  as  defined  in  Theorem  2.3,  is  known  as  the  flux  of  F  across  C. 
For  example,  if  F  represents  the  velocity  vector  field  of  a  planar  fluid,  then  the 
flux  measures  the  rate  of  fluid  transported  across  C  per  unit  time.  (We  assume 
that  F  does  not  vary  with  time  t.) 


6.2  I  Green's  Theorem  433 


To  see  this,  consider  the  amount  of  fluid  transported  across  a  small  segment 
of  C  during  a  brief  time  interval  At.  As  suggested  by  Figure  6.24,  we  have 


where  As  is  the  length  of  the  segment  of  the  curve  C.  Formula  (2)  is  only  ap- 
proximate, because  the  segment  of  curve  need  not  be  completely  straight  (so  the 
parallelogram  geometry  is  only  approximate),  and  because  the  vector  field  F  need 
not  be  constant  (so  the  term  F(x ,  y)  At  only  approximates  a  flow  line  of  F  over  the 
time  interval).  If  we  divide  formula  (2)  by  Af,  then  the  average  rate  of  transport 
across  the  segment  during  the  time  interval  Af  is  (F(x ,  y)  •  n)  As.  If  we  now  break 
up  C  into  finitely  many  such  small  segments,  sum  the  contributions  of  the  form 
(F(x ,  y)  •  n)  As  for  each  segment,  and  let  all  the  lengths  As  tend  to  zero,  we  find 
that  the  average  rate  of  fluid  transport,  denoted  A  M  /  A  t ,  is  given  approximately  by 


Finally,  letting  Af  — »•  0,  we  define  the  (instantaneous)  rate  of  transport  dM/dt 
to  be 


which  is  the  flux. 

In  view  of  the  remarks  above,  Theorem  2.3  tells  us  that  the  flux  of  F  across 
the  boundary  of  plane  region  D  (i.e.,  what  of  F  flows  across  dD)  is  equal  to 
the  total  (or  net)  divergence  of  F  over  all  of  D.  We  revisit  the  notion  of  flux  in 
Chapter  7  in  the  case  of  a  three-dimensional  vector  field  F.  In  that  setting,  we  are 
interested  in  the  flux  of  F  across  a  surface  in  R3 ;  defining  such  a  concept  requires 
a  surface  integral.  Then  Theorem  2.3  generalizes  to  three  dimensions  as  Gauss's 
theorem  (also  called  the  divergence  theorem).  Gauss's  theorem  relates  the  flux 
of  a  three-dimensional  vector  field  F  across  a  closed  surface  to  the  triple  integral 
of  the  divergence  of  F  over  the  solid  region  enclosed  by  the  surface. 

Proof  of  Green's  Theorem   

We  establish  Green's  theorem  (Theorem  2.1)  in  three  major  steps.  The  first  two 
steps  consist  of  proofs  of  special  cases  of  Theorem  2. 1 .  The  third  step  is  an  outline 
of  how  the  special  cases  may  be  used  to  provide  a  full  proof  of  the  general  case. 
As  a  result,  we  fall  short  of  a  complete,  rigorous  proof  of  the  very  general  version 
of  Green's  theorem  stated  in  Theorem  2.1.  However,  what  we  do  prove  makes  use 
of  the  important  geometric  ideas  of  multiple  integration  and  line  integrals,  and 
what  we  do  not  prove  is  rather  technical. 


Step  1.  We  establish  Green's  theorem  when  D  is  an  elementary  region  in  R2 
of  type  3.  Thus,  D  can  be  described  in  two  ways  (see  Figure  6.25): 

D  =  [(x,  y)  6  R2  |  y(x)  <  y  <  S(x),  a  <  x  <  b] 
=  {(x,  y)  6  R2  |  a(y)  <  x  <  P(y),  c<y  <d}. 


Amount  of  fluid  transported  %  area  of  parallelogram 

«  (F(x,  y)At  -n)As, 


(2) 


The  first  description  of  D  is  as  a  type  1  elementary  region;  the  second  is  as  a  type 
2  region.  (Recall  that  a  type  3  region  is  one  that  is  of  both  type  1  and  type  2.)  We 
assume  that  the  functions  a,  f),y,  and  S  are  all  continuous  and  piecewise  Cl. 


434       Chapter  6  |  Line  Integrals 


y  =  8(x) 


Figure  6.25  A  type  3  elementary  region  D  analyzed  as  both  a  type  1  and 
type  2  region.  Note  the  orientations  of  the  boundary  curves. 


Viewing  D  as  a  type  1  elementary  region,  we  evaluate  part  of  <faD  Mdx  + 
Ndy,  namely,  §aD  Mdx.  Note  that  3D  consists  of  a  lower  curve  C\  and  an  upper 
curve  Cj-  If  we  parametrize  these  curves  as  follows: 


x  =  t 


C\  :  \  > ,     a  <  t  <  b    and    C2  . 

}y  =  y(t)        ~    ~  \y  =  S(t) 


x  =  t 


a  <  t  < 


then  C2  is  oriented  opposite  to  the  desired  orientation  shown  in  Figure  6.25. 
Bearing  this  in  mind  we  compute 

M(x,y)dx  =  I   M{x,y)dx  —  I  M(x,y)dx. 
3D  JCi  Jc2 

(Note  the  minus  sign!)  Then 

(b   M(x,y)dx=      M(t,y(t))dt-  M(t,S(t))dt 

JdD  J  a  J  a 


-f 

J  a 


[M(t,y(t))~  M(t,S(t))]  dt. 


Now  we  compare  the  calculation  of  §dDMdx  with  that  of  / fD—(dM/dy)  dA. 
We  have 


b  rS(x) 


J  JD       dy  J  a  Jy 


dM 
~dy~ 


dy  dx 


-i 
-i 


[-M(x,  8(x))  +  M(x,  y(x))]dx 


[M(x,  y{x))  -  M(x,  8(x))]  dx. 


(Note  that  the  fundamental  theorem  of  calculus  was  used  here.)  Thus,  we  see  that, 
in  this  case, 


Mdx 


3D 


dM 

~dy~ 


dA. 


6.2  I  Green's  Theorem  435 


Figure  6.26  The  region 
fl  =  fliUD2Ufl3U  D4,  where 
each  subregion  D,-,  i  =  1,  2,  3,  4,  is 
elementary  of  type  3. 


Not  part 
of  dD 


Part  of  dD 

\ 


Part  of  dD 


Figure  6.27  The  common 
boundary  of  subregions  D;  and  Dj 
is  oriented  one  way  as  part  of  9  D, 
and  the  opposite  way  as  part  of 
dDj.  Hence,  fno,  Mdx  +  Ndy  + 
(fgD  Mdx  +  Ndy  will  cancel  over 
this  common  boundary. 


In  an  analogous  manner,  we  can  show 

f  f  dN 


dA 


dD 


by  viewing  D  as  a  type  2  elementary  region.  We  omit  the  details,  except  to  say 
that  both  the  line  integral  and  the  double  integral  can  be  shown  to  be  equal  to 


[N(p(y),y)-N(a(y),  y)]dy. 
Finally,  since  D  is  simultaneously  of  type  1  and  type  2, 


M(x,  y)dx  +  N(x,  y)dy  =  (b  Mdx 

3D  J3D 


dM 


Ndy 


dD 


Wd  3v 

-//. 


dA  + 


IL 


dN 
dx 


dA 


dN  dM\ 

 dA. 

dx       dy  ) 


Step  2.  Now  suppose  that  D  is  not  an  elementary  region  of  type  3,  but 
that  it  can  be  subdivided  into  finitely  many  type  3  regions  D\,  D2, . . . ,  Dn  in 
such  a  way  that  these  subregions  overlap  at  most  two  at  a  time  and  only  along 
common  boundaries.  Such  a  region  D  would  look  something  like  the  one  shown  in 
Figure  6.26.  By  Step  1 ,  Green's  theorem  holds  for  each  subregion.  Hence,  we  have 


fL(N- 


My)dA=  I J  (Nx  —  My)dA  +  /  /  (Nx-My)dA 

J  J Di  J  J  D2 

J j^{Nx-My)dA 


+  •••  + 


JdDi 


M  dx  +  N  dy 


M  dx  +  N  dy 


i 

J  3D,, 


H  h  <b     Mdx  +  Ndy. 


(3) 


At  this  point,  it  is  tempting  to  conclude  immediately  that  the  sum  of  the  line 
integrals  in  equation  (3)  is  <f8D  Mdx  +  Ndy.  However,  3D,  may  contain  more 
than  only  portions  of  dD.  The  trick  is  to  note  that  any  portion  of  3D,  that  is  not 
part  of  3D  is  part  of  exactly  one  other  3D7  .  Moreover,  this  overlapping  portion 
is  given  one  orientation  by  3D,  and  the  opposite  orientation  by  3D7-.  When  we 
take  the  sum  of  the  line  integrals  in  equation  (3),  any  contributions  arising  from 
the  components  of  the  3  D,-  's  that  are  in  the  interior  of  D  will  cancel  in  pairs. 
(See  Figure  6.27.)  Therefore, 


//.<* 


My)dA=(b    Mdx  +  Ndy+i)    Mdx  +  Ndy 

JdDi  J3D2 

H  h  <j>    Mdx  +  Ndy 

J  3D,, 


J  3D 


Mdx  +  Ndy; 


Green's  theorem  is  established  in  this  case. 


436       Chapter  6  |  Line  Integrals 


6.2  Exercises 


Step  3.  Unfortunately,  not  all  regions  described  in  the  statement  of  Theo- 
rem 2.1  can  be  subdivided  into  finitely  many  elementary  regions  of  type  3.  Here 
is  an  outline  of  what  we  might  do  to  prove  Green's  theorem  in  such  generality. 

First,  we  claim  (without  proof)  that  for  regions  D  described  in  the  statement 
of  Theorem  2. 1  we  can  produce  a  sequence  of  regions  D\ ,  D2, . . . ,  Dn , . . .  whose 
"limit"  as  n  — >  oo  is  D  and  such  that  each  Dn  can  be  subdivided  into  finitely  many 
type  3  elementary  regions.  Next,  we  claim  that  3D„  ->  3D  as  n  ->  oo.  Finally, 
we  need  to  prove  that,  as  n  — >•  oo, 

j  jD  (Nx  -  My)  dA     ->     J JjNx  -  My)  dA 

and 


Mdx  +  Ndy  <b    Mdx  +  Ndy. 

dD„  Jao 

Since  Green's  theorem  holds  for  each  Dn  (by  Steps  1  and  2),  we  are  done. 
Historical  Note3   


The  idea  that  the  line  integral  of  a  vector  field  along  a  closed  curve  can  be  related 
to  a  double  integral  over  the  region  bounded  by  the  curve  is  frequently  attributed 
to  George  Green  (1793-1841),  a  self-educated  English  mathematician.  The  result 
we  have  been  calling  Green's  theorem  had  its  origins  in  a  rather  obscure  1828 
pamphlet  published  by  Green,  in  which  he  sought  to  lay  a  rigorous  mathematical 
foundation  for  the  physics  of  electricity  and  magnetism.  Green's  ideas  arose  from 
work  in  partial  differential  equations  concerning  gravitational  potentials.  Green's 
pamphlet  subsequently  came  to  the  attention  of  Lord  Kelvin  (1824-1907),  who 
had  it  republished  so  that,  fortunately,  Green's  results  received  greater  recognition. 

Coincidentally,  a  result  similar  to  Green's  theorem  was  established  indepen- 
dently (and  also  in  1828!)  by  the  Russian  mathematician  Mikhail  Ostrogradsky 
(1801-1861).  Ostrogradsky 's  name  is  sometimes  associated  to  what  we  call 
Green's  theorem. 


In  Exercises  1—6,  verify  Green  s  theorem  for  the  given  vector  2.  F  =  (x2  —  y)  i  +  (x  +  y2)},    D    is    the  rectangle 

field  bounded  by  x  =  0,  x  =  2,  y  =  0,  and  y  =  1 . 

F  =  M(x,y)i  +  N(x,y))  3.  F  =  y i  +  x2 j,  D  is  the  square  with  vertices  (1,  1), 

(-1,  1),  (-1,-1),  and  (1,-1). 

and  region  D  by  calculating  both 

4.  F  =  2y  i  +  x  j,  D  is  the  semicircular  region  x2  +  y2  < 
(j)    Mdx  +  Ndy    and           (Nx  -  My)dA.  a2,y>0. 

JdD  J  JD  , 

5.  F  =  1y'\  —  Ax  j,  D  is  the  elliptical  region  x  + 
1 .  F  =  -x2y  i  +  xy2  j,  D  is  the  disk  x2  +  y2  <  4.  2y2  <  4. 


2  For  details  of  the  type  of  limit  argument  we  have  in  mind,  see  O.  D.  Kellogg,  Foundations  of  Potential 
Theory  (Springer,  Berlin,  1929;  reprinted  by  Dover  Publications,  New  York,  1954),  pp.  113-1 19,  where 
a  limit  argument  is  given  in  the  case  of  Gauss's  theorem,  which  we  explore  in  §7.3.  For  a  proof  of  Green's 
theorem  that  avoids  the  limit  argument,  see  D.  V  Widder,  Advanced  Calculus,  2nd  ed.,  (Prentice-Hall, 
Englewood  Cliffs,  1961;  reprinted  by  Dover  Publications,  New  York,  1989),  pp.  223-225. 

3  See  also  M.  Kline,  Mathematical  Thought  from  Ancient  to  Modern  Times  (Oxford  Press,  New  York, 
1972),  p.  683. 


6.2  |  Exercises  437 


6.  F  =  (x2y  +  x)i  +  (y3  —  xy2)j,  D  is  the  region  in- 
side the  circle  x2  +  y2  =  9  and  outside  the  circle 
x2  +  y2  =  4. 

7.  (a)  Use  Green's  theorem  to  calculate  the  line  integral 


y2dx  +  x2dy, 


where  C  is  the  path  formed  by  the  square  with 
vertices  (0,  0),  (1,  0),  (0,  1),  and  (1,  1),  oriented 
counterclockwise . 

(b)  Verify  your  answer  for  part  (a)  by  calculating  the 
line  integral  directly. 

8.  Let  F  =  2>xy  i  +  2x2  j  and  suppose  C  is  the  oriented 
curve  shown  in  Figure  6.28.  Evaluate 


F  •  ds 


both  directly  and  also  by  means  of  Green's  theorem. 


(2,-2) 


Figure  6.28  The  oriented  curve  C  of 
Exercise  8  consists  of  three  sides  of  a 
square  plus  a  semicircular  arc. 


9.  Evaluate 


(x2-y2)dx  +  (x2  +  y2)dy, 


where  C  is  the  boundary  of  the  square  with  vertices 
(0,  0),(1,0),  (0,  1),  and(l,  1),  oriented  clockwise.  Use 
whatever  method  of  evaluation  seems  appropriate. 

1 0.  Use  Green's  theorem  to  find  the  work  done  by  the  vec- 
tor field 

F  =  (4v-3x)i  +  (x-4y)j 

on  a  particle  as  the  particle  moves  counterclockwise 
once  around  the  ellipse  x2  +  Ay2  =  4. 

11.  Verify  that  the  area  of  the  rectangle  R  =  [0,a]  x  [0,  b] 
is  ab,  by  calculating  an  appropriate  line  integral. 


12.  Let  a  be  a  positive  constant.  Use  Green's  theorem  to 
calculate  the  area  under  one  arch  of  the  cycloid 


x  =  a(t  —  s'mt), 


a{\  —  cost). 


13.  Evaluate  <fc(x4y5  —  2y)dx  +  (3x  +  x5y4)dy,  where 
C  is  the  oriented  curve  pictured  in  Figure  6.29. 

y 

4 


Figure  6.29  The  oriented  curve  C  of 
Exercise  13. 

14.  Use  Green's  theorem  to  find  the  area  enclosed  by  the 
hypocycloid 

x(f)  =  (a  cos3  t,  a  sin3  t),     0  <  t  <  2n. 

15.  (a)  Sketch  the  curve  given  parametrically  by  x(f)  = 

(1  -t2,t3  -t). 

(b)  Find  the  area  inside  the  closed  loop  of  the  curve. 

16.  Use  Green's  theorem  to  find  the  area  between  the 
ellipse  x2/9  +  y2/4  =  1  and  the  circle  x2  +  y2  =  25. 

17.  Show  that  if  D  is  a  region  to  which  Green's  theorem 
applies,  and  3D  is  oriented  so  that  D  is  always  on  the 
left  as  we  travel  along  3D,  then  the  area  of  D  is  given 
by  either  of  the  following  two  line  integrals: 


Area  of  D 


x  dy 


y  dx. 


18.  Find  the  area  inside  the  quadrilateral  whose  vertices 
taken  counterclockwise  are  (2,  0),  (1,  2),  (—1,  1),  and 
(1,1). 

1 9.  Suppose  that  the  successive  vertices  of  an  n-sided  poly- 
gon are  the  points  (a\,  b\),  (02,  bi),  . . .  (an,  bn),  ar- 
ranged counterclockwise  around  the  polygon.  Show 
that  the  area  inside  the  polygon  is  given  by 


oi  b\ 

«2  bj 


+ 


a2  b2 
03  £>3 


+  •••  + 


«n-l  bn-1 

a„  b„ 


+ 


a„  b„ 
A]  bi 


20.  Let  a  be  a  positive  integer  throughout  this  problem. 
An  epicycloid  is  the  path  produced  by  a  marked  point 
on  a  circle  of  unit  radius  that  rolls,  without  slipping, 


438       Chapter  6  |  Line  Integrals 


on  the  outside  of  a  fixed  circle  of  radius  a.  If  the  cen- 
ter of  the  fixed  circle  is  at  the  origin  and  the  marked 
point  is  at  (a,  0)  when  t  =  0,  then  the  epicycloid  is 
given  by  the  path  \(t)  =  ((a  +  1)  cos  t  —  cos  (a  +  1  )f , 
(a  +  l)sin?  —  sin(a  +  l)r).  (See  Exercise  35  of  the 
Miscellaneous  Exercises  for  Chapter  1.) 

(a)  Show  that  the  epicycloid  path  meets  the  fixed  cir- 
cle exactly  when  t  =  Inn /a,  where  n  is  an  integer. 
(Hint:  This  must  happen  when  \\x(t)\\  =  a.)  Graph 
the  epicycloid  when  a  =  5,6. 

(b)  Use  an  appropriate  line  integral  to  find  the  area 
enclosed  by  the  epicycloid. 

(c)  As  the  integer  a  gets  larger,  what  happens  to  the 
ratio  of  the  area  calculated  in  part  (b)  to  that  of  the 
fixed  circle? 

21.  Evaluate  the  line  integral  §c  5y  dx  —  3x  dy,  where  C 
is  the  cardioid  with  polar  coordinate  equation  r  = 
1  —  sin  6,  oriented  counterclockwise. 

22.  (a)  Suppose  that  C  is  a  simple,  closed  curve  that  does 

not  enclose  the  origin.  Use  Green's  theorem  to  de- 
termine the  value  of 

x  dx  +  y  dy 


ic  xz  +  r 

(b)  Now  suppose  that  C  is  a  simple,  closed  curve  that 
does  enclose  the  origin.  Can  you  use  Green's  the- 
orem to  determine  the  value  of 

x  dx  +  y  dy 
x2  +  y2 
Explain. 

(c)  Let  C\  and  C2  be  two  simple,  closed  curves  that 
both  enclose  the  origin,  are  both  oriented  coun- 
terclockwise, and  do  not  touch  or  intersect.  Show 
that 

x  dx  +  y  dy       f  x  dx  +  y  dy 


X1-  +  yl        JCl     x^  +  y^ 
(d)  Use  the  result  of  part  (c)  to  determine  the  value  of 

x  dx  +  y  dy 


ic  *2  +  r 

where  C  is  a  simple,  closed  curve  that  encloses  the 
origin. 

23.  (a)  Use  the  divergence  theorem  (Theorem  2.3)  to  show 

that  <fc  F  •  rids  =  0,  where  F  =  2y  i  —  3x  j  and  C 
is  the  circle  x2  +  y2  =  I. 

(b)  Now  show  (fc  F  •  n  ds  =  0  by  direct  computation 
of  the  line  integral. 

24.  Let  F  =  M(x,  y)\  +  N(x,  y) j.  The  divergence  theo- 
rem shows  that  the  flux  of  F  across  a  closed  curve  C 


(i.e.,  <fc  F  -nds)  is  equal  to  / fD(Mx  +  Ny)  dA,  where 
D  is  the  region  bounded  by  C.  Use  Green's  theorem 
to  establish  a  similar  result  involving  (fc  F  •  Tds,  the 
circulation  of  F  along  C.  (See  also  §6.1.) 

25.  Let  C  be  any  simple,  closed  curve  in  the  plane.  Show 
that 


26.  Show  that 


3x  y  dx  +  x  dy  =  0. 


-y3  dx  +  (jc3  +  2x  +  y)dy 


is  positive  for  any  closed  curve  C  to  which  Green's 
theorem  applies. 

27.  Show  that  if  C  is  the  boundary  of  any  rectangular  re- 
gion in  R2,  then 


3y)  dx  +  x3y2  dy 


depends  only  on  the  area  of  the  rectangle,  not  on  its 
placement  in  R2 . 

28.  Let  r  =  x  i  +  y  j  be  the  position  vector  of  any  point 
in  the  plane.  Show  that  the  flux  of  F  =  r  across  any 
simple  closed  curve  C  in  R2  is  twice  the  area  inside  C. 

29.  Let  D  be  a  region  to  which  Green's  theorem  applies 
and  suppose  that  u(x,  y)  and  v(x,  y)  are  two  functions 
of  class  C2  whose  domains  include  D.  Show  that 


3(m,  v) 


dA 


(mVd)  •  ds, 


where  C  =  3D  is  oriented  as  in  Green's  theorem. 
30.  Let  f(x,  y)  be  a  function  of  class  C2  such  that 


a2/  a2/ 

dx2  9v2 


0 


(i.e.,  /  is  harmonic).  Show  that  if  C  is  any  closed  curve 
to  which  Green's  theorem  applies,  then 


9/ 
By 


dx 


dx 


dy  =  0. 


31.  Let  D  be  a  region  to  which  Green's  theorem  applies 
and  n  the  outward  unit  normal  vector  to  D.  Suppose 
f(x,  y)  is  a  function  of  class  C2.  Show  that 

3/ 


f  f  V2fdA=(f 

J  J D  JSL 


9n 


ds, 


where  V2/  denotes  the  Laplacian  of  /  (namely, 
V2/  =  d2f/dx2  +  92//3y2)  and  df/dn  denotes 
V/  •  n.  (See  the  proof  of  Theorem  2.3  for  more  in- 
formation about  n.) 


6.3  |  Conservative  Vector  Fields  439 


6.3  Conservative  Vector  Fields 


Figure  6.30  If  F  has  path- 
independent  line  integrals,  then 
fc  F  •  ds  =  fc  F  •  ds  for  any  two 
piecewise  C1  oriented  curves  from 
A  to  B. 


As  seen  in  §6. 1 ,  line  integrals  of  a  given  vector  field  depend  only  on  the  underlying 
curve  and  its  orientation,  not  on  the  particular  parametrization  of  the  curve.  In 
some  special  instances,  however,  even  the  curve  itself  doesn't  matter,  only  the 
initial  and  terminal  points.  A  vector  field  having  the  property  that  line  integrals  of 
it  depend  only  on  the  initial  and  terminal  points  of  the  oriented  curve  over  which 
the  line  integral  is  taken  is  said  to  have  path-independent  line  integrals.  We  next 
state  a  more  careful  definition,  characterize  such  vector  fields,  and  explore  their 
significance. 

Path  Independence   


DEFINITION  3.1  A  continuous  vector  field  F  has  path-independent  line 
integrals  if 

/  F-ds  =  I  F-ds 

JCi  Jc2 

for  any  two  simple,  piecewise  C\  oriented  curves  lying  in  the  domain  of  F 
and  having  the  same  initial  and  terminal  points.  (See  Figure  6.30.) 


(1,1) 


Figure  6.31  The  curves 
Ci  and  Ci  of  Example  1. 


EXAMPLE  1  Let  F  =  y  i  —  x  j  and  consider  the  following  two  curves  in  R2 
from  the  origin  to  (1,  1):  C\,  the  line  segment  from  (0,  0)  to  (1,  1),  and  C%,  the 
portion  of  the  parabola  y  =  x2  (the  curves  are  shown  in  Figure  6.31).  These  curves 
may  be  parametrized  as 


:  \X     1     0  <  t  <  1    and  C2 


[y  =  t 
Then  we  calculate 


x  =  t 

y  =  t2 


o  <  t  <  1. 


f  F-ds=  I 

JC\  Jo 


(H-fj)-(i  +  j)df 


Jo 


Odt  =  0, 


while 


f  F-ds=  f  (/2i-/j).(i  +  2fj 

Jc2  Jo 


i)dt 


f 

Jo 


(t2-2t2)dt  = 


3'  lo 


We  see  that 


/   F-ds^  j  F-ds, 

JCi  Jc2 

and  so  F  does  not  have  path-independent  line  integrals. 


EXAMPLE  2  Let  F  =  xi  +  yj.  This  vector  field  has  path-independent  line 
integrals — we  will  see  why  presently.  For  the  moment,  however,  we  illustrate  (not 
prove)  this  fact  by  considering  the  parabolic  path  x:  [0,  1]  ->  R2,  x(f)  =  (t,  t2), 
as  well  as  the  path  y:  [0,2]  ->  R2  made  up  of  the  two  straight  segments 

y,:[0,  1]^R2,  yi(0  =  (0,0    and   y2:  [1,  2]  ->  R2,  y2(r)  =  (t  -  1,  1). 


440       Chapter  6  |  Line  Integrals 


Figure  6.32  The  paths  x 
and  y  (consisting  of  the 
paths  yi  and  y2)  of 
Example  2. 


Both  x  and  y  are  paths  from  (0,  0)  to  (1 ,  1)  and  are  shown  in  Figure  6.32.  We  have 
Jv-ds  =  j  (t\  +  t2])-(\  +  2t\)dt 

=  f  (t  +  2t3)dt=  {\t2  +  \tA)\[=  1, 
Jo 


and 


F-ds  =  /  F-ds  +  /  F-ds 

Jy  Jyi  Jy2 


j  fj'j*  +  ^  (ft  -  1)1  + JM* 


(t  -  \)dt 


2'    lo  ^  2^'       L>  \\        2    1  2 


1. 


To  establish  that  F  has  path-independent  line  integrals,  we  would  need  to  check 
that  the  value  of  the  line  integral  of  F  along  any  choice  of  path  between  any  two 
points  is  the  same  as  any  other — a  prohibitive  task.  ♦ 


The  following  result  is  a  reformulation  of  the  path-independence  property: 


THEOREM  3.2  Let  F  be  a  continuous  vector  field.  Then  F  has  path-independent 
line  integrals  if  and  only  if  §CF  -ds  =  0  for  all  piecewise  C1,  simple,  closed 
curves  C  in  the  domain  of  F. 


PROOF  First,  assume  that  F  satisfies  the  path-independence  property.  Suppose 
C  is  parametrized  by  x(t),  a  <  t  <  b,  where  x(a)  =  x(b)  =  A.  Let  B  be  another 
point  on  C,  and  break  C  into  two  oriented  curves  C\  and  C2  from  A  to  B.  One 
of  these  curves — say,  C\ — will  be  oriented  the  same  way  as  C  and  the  other  the 
opposite  way.  (See  Figure  6.33.)  Thus, 

Figure  6.33  The  simple,  (f)  F  •  ds  =  f  F  •  ds  -  f  F-Js  =  0, 

closed  curve  C  consists  of  Jc  JC{  JCl 

two  oriented  curves  from  A 

t0  5  since  F  has  path-independent  line  integrals. 

Conversely,  suppose  that  all  line  integrals  of  F  around  simple,  closed  curves 
are  zero.  Then  given  two  piecewise  C\  oriented,  simple  curves  C\  and  C2  with 
the  same  initial  and  terminal  points,  let  C  be  the  closed  curve  consisting  of  C\ 
and  — C2  (i.e.,  C2  with  its  direction  reversed).  Then,  we  have 


(t)F-ds=      F-ds+        F-ds=      F-ds-  F-ds. 

JC  JCi  J-C2  JCi  Jc2 

If  C  happens  to  be  simple,  then  j>c  F  -  ds  =  0  by  assumption,  so 

/   F-ds=  F-ds, 

Jc,  JCi 


6.3  |  Conservative  Vector  Fields  441 


as  desired.  However,  C  need  not  be  simple  even  if  C\  and  C2  are.  (See  Figure  6.34.) 
If  it  is  possible  to  break  C\  and  C2  into  finitely  many  segments  so  that  a  segment 
C[  of  C\  and  a  segment  C'2  of  C2  either  (i)  completely  coincide  or  (ii)  together 
form  a  simple,  closed  curve  C,  then  it  is  not  too  difficult  to  modify  the  preceding 
argument  to  conclude  that 


Figure  6.34  The  closed  curve  C 

constructed  from  C\  and  the 

reverse  of  C2  need  not  be  simple. 

However,  it  is  not  always  possible  to  do  this,  and  further  technical  arguments 
(which  we  omit)  are  required.  ■ 

We  remark  that  it  is  not  essential  to  assume  that  the  curves  in  Definition  3.1 
and  Theorem  3.2  are  simple.  We  have  done  so  here  in  order  to  make  the  proof  of 
Theorem  3.5  below  more  straightforward. 


/  F-ds  =      F  •  ds. 

Jc,  Jc2 


Gradient  Fields  and  Line  Integrals   

We  describe  next  a  class  of  vector  fields  that  satisfy  the  path-independence  prop- 
erty, namely,  gradient  fields.  Suppose  that  F  is  a  continuous  vector  field  such  that 
F  =  V/,  where  /  is  some  scalar-valued  function  of  class  Cl.  (Recall  that  we 
refer  to  /  as  a  scalar  potential  of  F.  We  also  call  F  a  conservative  vector  field 
as  well  as  a  gradient  field).  Then,  along  any  path  x  from  A  =  x(a)  to  B  =  x(b) 
whose  image  lies  in  the  domain  of  F,  we  have 

/"f-Js=  fvf-ds=  f  Wf(x(t))-x'(t)dt. 

Jx  Jx  J  a 

It  follows  from  the  chain  rule  that  d/dt[f(x(t))]  =  V/(x(r))  •  x'(0-  Hence, 
fjf-d*  =  f  V/(x(0)  •  AO  dt  =  fjt  dt 

=  f(x(t))\ha  =  f(x(b))  -  f(x(a))  =  f(B)  -  f(A). 

Therefore,  when  F  is  a  gradient  field  the  line  integral  of  F  depends  only  on  the 
value  of  the  potential  function  at  the  endpoints  of  the  path.  Hence,  gradient  fields 
have  path-independent  line  integrals.  The  converse  holds  as  well,  as  we  prove  at 
the  end  of  this  section.  Stated  formally,  we  have  the  following  theorem: 


THEOREM  3.3  Let  F  be  defined  and  continuous  on  a  connected,  open  region  R 
of  R".  Then  F  =  V/  (where  /  is  a  function  of  class  C1  on  R)  if  and  only  if  F  has 
path-independent  line  integrals  over  curves  in  R.  Moreover,  if  C  is  any  piecewise 
C  ,  oriented  curve  lying  in  R  with  initial  point  A  and  terminal  point  B,  then 

jv.di  =  f(B)-f(A). 


Note:  A  region  R  C  R"  is  connected  if  any  two  points  in  R  can  be  joined  by 
a  path  whose  image  lies  in  R. 

EXAMPLE  3  Consider  the  vector  field  F  =  x  i  +  y  j  of  Example  2  again.  You 
can  readily  check  that  F  =  V/,  where  f(x,  y)  =  \{x2  +  y2).  By  Theorem  3.3, 
line  integrals  of  F  will  be  path  independent;  this  fact  was  illustrated,  but  not 
proved,  in  Example  2  when  we  calculated  the  vector  line  integral  of  F  along  two 


442       Chapter  6  |  Line  Integrals 


paths  from  (0,  0)  to  (1,  1).  By  Theorem  3.3,  we  see  now  that  for  any  directed 
piecewise  C1  curve  C  from  (0,  0)  to  (1,  1),  we  have 

J^F-ds  =  f(l,  1)—  /(0,  0)  =  \{\2  +  l2)  -  i(02  +  02)  =  1, 

which  agrees  with  our  earlier  computations.  ♦ 


A  Criterion  for  Conservative  Vector  Fields  — 

Theorem  3.3  tells  us  that  a  vector  field  F  has  path-independent  line  integrals 
precisely  when  it  is  a  conservative  (gradient)  vector  field  and  moreover,  that  the 
line  integral  of  F  along  any  path  is  determined  by  the  values  of  the  potential 
function  /  at  the  endpoints  of  the  path.  Two  questions  arise  naturally: 

1.  How  can  we  determine  whether  a  given  vector  field  F  is  conservative? 

2.  Assuming  that  F  is  conservative,  is  there  a  procedure  for  finding  a  scalar 
potential  function  /  such  that  F  =  V/? 

We  answer  the  first  question  by  providing  a  simple  and  effective  test  that  can 
be  performed  on  F.  Should  F  pass  this  test  (i.e.,  if  F  is  conservative),  then  we 
illustrate  via  examples  how  to  produce  a  scalar  potential  for  F,  thereby  answering 
the  second  question. 

First,  we  need  additional  terminology. 


DEFINITION  3.4  A  region  R  in  R2  or  R3  is  simply-connected  if  it  con- 
sists of  a  single  connected  piece  and  if  every  simple,  closed  curve  C  in  R 
can  be  continuously  shrunk  to  a  point  while  remaining  in  R  throughout  the 
deformation. 


If  R  is  a  region  in  the  plane,  then  R  is  simply-connected  just  in  case  it  is 
connected  and  every  simple,  closed  curve  C  lying  in  R  has  the  property  that  all 
the  points  enclosed  by  C  also  lie  in  R.  Loosely  speaking,  a  simply-connected 
region  (in  either  R2  or  R3)  can  have  no  "essential  holes."  Illustrative  examples  are 
shown  in  Figures  6.35  and  6.36.  The  notion  of  continuously  shrinking  a  curve  to  a 
point  can  be  made  fully  precise,  although  we  shall  not  take  the  trouble  to  do  so  here. 


(1)  (2) 

Figure  6.35  (1)  The  region  Ri  C  R2  is  simply-connected:  All  points 
surrounded  by  any  simple,  closed  curve  in  Ri  lie  in  R i .  (2)  In  contrast,  R2  is  not 
simply-connected:  Although  the  curve  C\  encloses  points  that  lie  in  R2,  the 
curve  C2  surrounds  a  hole.  Hence,  C2  cannot  be  continuously  shrunk  to  a  point 
while  remaining  in  R2 . 


6.3  I  Conservative  Vector  Fields  443 


(2) 


c 


{(0,0,0)) 


R7  =  & 


■  z-axis 


Figure  6.36  (1)  The  region  R\  C  R3  is  simply-connected.  (2)  The  region  R2  is  not 
simply-connected:  The  curve  C  cannot  be  shrunk  continuously  to  a  point  without 
becoming  "stuck"  on  the  "missing"  z-axis. 

Now  we  state  our  criterion  for  a  vector  field  to  be  conservative. 

THEOREM  3.5  Let  F  be  a  vector  field  of  class  C1  whose  domain  is  a  simply- 
connected  region  R  in  either  R2  or  R3.  Then  F  =  V/  for  some  scalar- valued 
function  /  of  class  C2  on  R  if  and  only  if  V  x  F  =  0  at  all  points  of  R. 


Before  proving  Theorem  3.5,  some  remarks  and  examples  are  appropriate. 
First,  note  that  Theorem  3.5  provides  a  straightforward  way  to  determine  if  a  vector 
field  F  is  conservative:  Check  that  the  domain  of  F  is  simply-connected  and  then 
test  if  V  x  F  =  0.  If  the  curl  vanishes,  it  follows  that  F  has  path- independent  line 
integrals.  This  "curl  criterion"  can  be  helpful  in  practice. 

In  the  case  where  F  =  M(x ,  y)  i  +  N(x ,  y)  j  is  a  two-dimensional  vector  field 
the  condition  that  the  curl  of  F  vanishes  means 


dN 
~dx~ 


i  i  k 

d/dx         d/dy  d/dz 
M(x,y)     N(x,y)  0 

This  is  equivalent  to  the  condition 

dN  _  dM 

dx  dy 

Equation  (1)  is  a  simpler  condition  to  use  in  this  situation. 


dM\ 

  k  =  0. 

dy  J 


(1) 


EXAMPLE  4   LetF  =  x2yi-2xyj.Then 


dx 


(— 2xy)  =  —2y  and 


dy 


(x2y)  =  x2. 


Since  these  partial  derivatives  are  not  equal,  we  conclude  that  F  is  not  conservative, 
by  Theorem  3.5.  ♦ 

EXAMPLE  5  Let  F  =  (2xy  +  cos 2y)i  +  (x2  -  2x  sin2y)j.  The  vector  field 
F  is  defined  and  of  class  C1  on  all  of  R2  (a  simply-connected  region).  Moreover, 


— (x  —  2x  sin2y)  =  2x  —  2  sin2y   and   — (2xy  +  cos2y)  =  2x  —  2  sin2y. 

dx  dy 


444       Chapter  6  I  Line  Integrals 


We  may  conclude  that  F  is  conservative.  In  addition,  if  C  is  the  ellipse 
x2/4  +  v2  =  1  (a  simple,  closed  curve),  then,  by  Theorems  3.2  and  3.3,  we  con- 
clude, without  any  explicit  calculation,  that  j>c  F  •  ds  =  0.  ♦ 

EXAMPLE  6  Let 


F  = 


x1  +  y2  +  z2 


6x   i  + 


x 2  +  yz  +  z 


+  y  +  Z 


F  is  of  class  C1  on  all  of  R3  except  for  the  origin.  Note  that  R3  —  {(0,  0,  0)}  is 
simply-connected.  We  leave  it  to  you  to  check  that  V  x  F  =  0  for  all  (x,  y,  z)  in 
the  domain  of  F.  Therefore,  by  Theorem  3.5,  F  is  conservative. 

Now  suppose  x:  [0,  1]  -»  R3  is  the  path  given  by  x(t)  =  (1  —  t,  shur/,  t).  To 
evaluate  fx  F  •  ds  directly,  we  must  calculate 


i 


(1  -  t)2  +  sin2  nt  +  t2 


6(1-0, 


sm7rf 


(1  -t)2  +  sm27tt  +  t 


2  ' 


(1,0,0) 


Figure  6.37  The  paths 

\(t)  =  (1  -  t,  sinjrf,  t),  0  <  t  <  1 

and  y(f )  =  (cos  r,  0,  sin  t ), 

0  <  ?  <  jt/2. 


(1  -  02  +  sin27rf  +  f: 

2r  —  1  +  it  sin itt  cositt 
(1  -  t)2  +  sin2  7if  +  t2 


(— 1,  7T  COS  7Tf ,  1)  c/f 


+  6(1  -  01* 


This  last  integral  is  tricky  to  evaluate.  However,  since  F  is  conservative,  we  may 
evaluate  F  by  calculating  /  F  •  ds,  where  y  is  any  other  path  with  the  same 
endpoints  as  x.  A  good  choice  is  y(0  =  (cos?,  0,  sinO,  0  <  t  <  n/2,  because 
the  image  of  this  path  lies  on  the  sphere  x2  +  y2  +  z2  =  1 ,  a  fact  that  will  en- 
able us  to  work  with  a  simple  integral.  (See  Figure  6.37  for  a  graph  of  the  two 
paths  x  and  y.)  Since  F  is  conservative  (and  hence  has  path-independent  line 
integrals), 


6cosf, 


0  sinf 


1  1 


(—  sinf,  0,  cos  t)dt 


-L 


n/2 


6  cos  t  sin  t  dt 


-L 


n/2 


3  s\n2t  dt 


-|cos2r|0 


n/2 


K-i-i)  =  3. 


Sketch  of  a  proof  of  Theorem  3.5  By  Theorem  4.3  of  Chapter  3,  note  that,  if 
F  =  V/  for  some  function  /  of  class  C2,  then  V  x  F  =  V  x  (V/)  =  0. 

Conversely,  suppose  that  V  x  F  =  0.  We  show  that  if  C  is  any  piecewise  C1, 
simple,  closed  curve  in  R,  then  §c  F  •  ds  =  0.  By  Theorem  3.2,  this  implies  that 
F  has  path-independent  line  integrals,  which,  by  Theorem  3.3,  is  equivalent  to 
F's  being  the  gradient  of  some  scalar- valued  function  /.  Moreover,  since  F  is 
assumed  to  be  of  class  C1,  it  follows  that  /  must  be  of  class  C2. 


6.3  |  Conservative  Vector  Fields  445 


Figure  6.38  Since  R  is 
simply-connected,  any  region  D 
enclosed  by  C  must  lie  in  R. 


Figure  6.39  A 

surface  in  space, 
bounded  by  the 
simple,  closed 
curve  C. 


To  see  that  §c^  •  ds  =  0,  suppose,  first,  that  F  is  defined  on  a  simply- 
connected  region  R  in  R2.  Since  R  is  simply-connected,  the  closed  curve  C 
bounds  a  region  D  that  is  entirely  contained  in  R.  (See  Figure  6.38.)  By 
Proposition  2.2,  which  is  equivalent  to  Green's  theorem,  we  have 


(f  ¥-ds  =  ±  f  f  (VxF-k)JA  =  ±  /  / 
Jc  J  Jd  J  Jd 


OdA  =  0. 


We  use  the  "±"  sign  in  the  event  that  C  is  oriented  opposite  to  the  orientation 
stipulated  by  Green's  theorem.  If  F  is  defined  on  a  simply-connected  region  R 
in  R3,  then  we  must  apply  Stokes's  theorem  rather  than  Green's  theorem.  This 
gets  us  a  little  ahead  of  ourselves,  although  the  principle  remains  the  same:  If  C 
is  a  simple,  closed  curve  in  R  C  R3,  then  <fc  F  •  ds  is  equal  to  a  suitable  surface 
integral  of  V  x  F  over  a  surface  in  R  bounded  by  C.  (See  Figure  6.39.)  Because 
the  curl  is  assumed  to  be  zero,  any  integral  of  it  will  be  zero,  and  so  §c  F  •  ds  is 
zero  as  well.  ■ 

Finding  Scalar  Potentials   

Now  that  we  have  a  practical  test  to  determine  whether  a  given  vector  field  F 
is  conservative,  we  illustrate  in  Examples  7  and  8  a  straightforward  method  for 
producing  a  scalar  potential  function  for  F.  This  technique  is  a  direct  consequence 
of  the  definition  of  a  gradient  field. 

EXAMPLE  7   Consider  the  vector  field 

F  =  (2xy  +  cos  2y)  i  +  (x2  —  2x  sin  2  y )  j 

of  Example  5.  We  have  already  seen  that  F  is  conservative  in  Example  5.  To  find 
a  scalar  potential  for  F,  we  seek  a  suitable  function  f(x,  y)  such  that 

V/(*,y)  =  —  i+  —  J  =  F. 
ox  ay 

The  components  of  the  gradient  of  /  must  agree  with  those  of  F;  therefore, 

Bf 
dx 

Bf 
3y 

We  may  begin  to  recover  /  by  integrating  the  first  equation  of  (2)  with  respect  to 
x.  Thus, 

fdff  7 
f(x,y)=  I  — dx  =  /  (2xy  +  cos2y)dx  =  x  y  +  x  cos2y  +  g(y),  (3) 
J  dx  J 

where  g(y)  is  an  arbitrary  function  of  y.  (The  function  g(y)  plays  the  role  of  the 
arbitrary  "constant  of  integration"  in  the  indefinite  integral  of  df/dx.)  Differen- 
tiating equation  (3)  with  respect  to  y  yields 

9/ 
dy 

If  we  compare  equation  (4)  with  the  second  equation  of  (2),  we  see  that  g'(y)  =  0, 
and  so  g  must  be  a  constant  function.  Therefore,  our  scalar  potential  must  be  of 
the  form 

f(x,  y)  =  x2y  +  x  cos2y  +  C, 


=  2xy  +  cos2y 
=  x2  —  2x  sin2y 


(2) 


2x  sin2y  +  g'(y). 


(4) 


446       Chapter  6  |  Line  Integrals 


where  C  is  an  arbitrary  constant.  You  may,  if  you  wish,  double-check  that 
V/  =  F.  ♦ 

EXAMPLE  8  LetF  =  (ex  siny  -  yz)i+  (ex  cosy  -  xz)\  +  (z  -  xy)k.Note 
that  F  is  of  class  C1  on  all  of  R3.  We  calculate 


i  j  k 

d/dx  d/dy  d/dz 

ex  sin  y  —  yz      ex  cos  y  —  xz      z  —  xy 


3  3 

—  (z-  xy)  -  —  (ex  cos  y  -  xz) 
ay  dz 


3  3 
+  (  7^(ex  siny  -  yz)  -  —  (z  -  xy) 


3  3 
(  g^(ex  cosy  -  xz)  -  —(ex  siny  -  yz) 


=  0. 

Therefore,  by  Theorem  3.5,  F  is  conservative. 

Any  scalar  potential  f(x,  y,  z)  for  F  must  satisfy 


V 

dx 

df 
8y 

df 
dz 


=  e  sin  y  —  yz 
=  ex  cos  y  —  xz- 
=  z  —  xy 


Integrating  df/dx  with  respect  to  x,  we  find  that 

^  dx 


=  j  (ex  sin  y  —  yz)  dx 


=  ex  siny  -  xyz  +  g(y,  z), 


(5) 


(6) 


where  g(y,  z)  may  be  any  function  of  y  and  z.  Differentiating  equation  (6)  with 
respect  to  y  and  comparing  with  the  second  equation  in  (5),  we  see  that 

M     x  9s  , 

—  =  e  cos  v  —  xz  H  =  e  cos  y  —  xz. 

dy  3y 

Hence,  dg/dy  =  0,  so  g  must  be  independent  of  y;  that  is,  g(y,  z)  =  h(z),  a 
function  of  z  alone.  So 


f(x,  y,z)  =  ex  siny  -  xyz  +  h(z). 


(7) 


6.3  |  Conservative  Vector  Fields  447 


Finally,  we  differentiate  equation  (7)  with  respect  to  z  and  compare  with  the  third 
equation  of  (5): 

—  =  -xy  +  h(z)  =  z  -  xy. 

dz 

Therefore,  h'(z)  =  z,  so  h(z)  =  \z2  +  C,  where  C  is  an  arbitrary  constant.  Thus, 
a  scalar  potential  for  the  original  vector  field  F  is  given  by 

f{x,  y,  z)  =  ex  sin  y  -  xyz  +  \z2  +  C.  + 


Addendum:  Proof  of  Theorem  3.3   

Recall  that  we  have  already  shown  that  if  a  vector  field  F  is  a  gradient  field  then  F 
has  path-independent  line  integrals.  So  now  we  need  only  establish  the  converse. 
We  do  this  explicitly  in  the  case  where  F  is  defined  on  a  (connected)  subset  R  of 
R3 ,  although  our  proof  requires  only  notational  modification  in  the  n  -dimensional 
setting. 

Assume  that  F  has  path-independent  line  integrals.  Then  we  may  unambigu- 
ously write  F  •  ds  to  denote  the  vector  line  integral  of  F  from  the  point  A 
to  the  point  B  along  any  path  whose  image  lies  in  R.  In  what  follows,  consider 
A(xo,  yo,  Zo)  to  be  a  fixed  point  in  R,  and  5(x,  y,  z)  a  "variable  point."  Then  we 
define 


KB) 


J  A 


ds 


and  show  that  /  is  a  scalar  potential  for  F. 
Write  F  explicitly  as 

F  =  Mix,  y,z)i  +  Nix,  y,  z)j  +  P(x,  y,  z)k. 

Therefore,  we  need  to  verify  that  the  components  of  V/  agree  with  those  of  F; 
that  is, 


Figure  6.40  If  B'  is 

sufficiently  close  to  B,  then  the 
straight-line  path  from  B  to  B' 
will  lie  inside  R. 


df 

dx 


=  Mix,  y,  z), 


df 

dy 


=  Nix,  y,  z),  and 


Bf 
dz 


=  Pix,y,z). 


Actually,  we  check  only  the  first  of  these  equations;  the  others  may  be  verified  in 
a  similar  manner. 

At  the  point  B,  we  have,  by  the  definition  of  the  partial  derivative,  that 


3/  ,.  fjx  +  h,y,z)-  fjx,y,z)  fjB')  -  fjB) 
—  =  hm  =  lim  , 

dx     h-+o  h  h-+o  h 


where  B'  denotes  the  point  (x  +  h,  y,  z).  Note  that  B' 
merator  of  the  difference  quotient  in  equation  (8)  is 


B  as  h 


fiB')-fiB)=  f  F-ds-  fBF-ds  =  ^  F 

J  A  J  A  JB 


ds. 


(8) 

0.  The  nu- 
(9) 


If  h  is  sufficiently  small,  we  may  evaluate  the  line  integral  in  equation  (9)  by  using 
the  straight-line  path  between  B  and  B' .  (See  Figure  6.40.)  Explicitly,  this  path  is 

x(t)  =  ix  +  th,  y,  z),     0  <  t  <  1. 


448       Chapter  6  |  Line  Integrals 

Then 


6.3  Exercises 


pB'  pi  pi 

/    F-ds  =      F(x  +  th,  y,z)-(h,0,  0)dt=  /   hM(x  +  th,  y,  z)dt. 
Jb  Jo  Jo 

Since  t  is  between  0  and  1,  \th\  <  \h\.  By  the  continuity  of  F  (and  therefore  M) 
we  have,  for  h  s»  0, 

M(x  +  th,  y,  z)  ~  M{x,  y,  z). 
This  approximation  improves  as  h  — »•  0.  Hence, 

f(B')-f(B)=  f    F-ds  «  /  hM(x,y,z)dt  =  hM(x,y,z). 
Jb  Jo 

Using  this  result  in  equation  (8),  we  see  that 


^  =  lim  ^-[/iM(x,  y,  z)]  =  M(x,  y,  z), 
dx      h^o  h 


as  desired. 


1.  Consider  the  line  integral  Jc  z2  dx  +  2y  dy  +  xz  dz. 

(a)  Evaluate  this  integral,  where  C  is  the  line  segment 
from  (0,0,  0)  to  (1,  1,  1). 

(b)  Evaluate  this  integral,  where  C  is  the  path 
from  (0,0,0)  to  (1,1,1)  parametrized  by 
x(f)  =  (r,  ?2,?3),  0  <  t  <  1. 

(c)  Is  the  vector  field  F  =  z2i  +  2yj+xzk  conser- 
vative? Why  or  why  not? 

2.  Let  F  =  2xy  i  +  (x2  +  z2)  j  +  2yz  k. 

(a)  Calculate  fc  F  •  ds,  where  C  is  the  path  param- 
etrized by  x(f)  =  (t2,  f3,  f5),  0  <  t  <  1. 

(b)  Calculate  Jc  F  •  ds,  where  C  is  the  straight-line 
path  from  (0,  0,  0)  to  (1,  0,  0),  followed  by  the 
straight-line  path  from  (1,  0,  0)  to  (1,  1,  1). 

(c)  Does  F  have  path-independent  line  integrals?  Ex- 
plain your  answer. 

In  Exercises  3—17,  determine  whether  the  given  vector  field 
F  is  conservative.  If  it  is,  find  a  scalar  potential  function  for  F. 

3.  F  =  ex+y  i  +  exy  j 

4.  F  =  2x  sin  y  i  +  x2  cos  y  j 

5.  F  =  (3a-2  cos  y  +  -    J—.  ,  )  i 

V  l+x2y2/ 

2  2 
XV  XV 

6.  F  =  i  H  — 

(1+x2)2  l+x2' 

7.  F  =  (e~y  —  y  sinxy)i  —  (ic"1'  +  x  sinxy)j 


8.  F  =  (6xy2  +  2y3)i  +  (6x2y-xy)j 

9.  F  =  (6xy2  -  3x2)i  +  (y2  +  6x2y)j 

10.  F  =  (xyz3  +  xy  -  z2)  i  +  (2x2z3  -  y2  +  2yz)  j 

+  (6x2y  -  y2z)k 

11.  F  =  (4xyz3  —  2xy)  i  +  (2x2z3  -  x2  +  2y z)  j 

+  (6x2yz2  +  y2)k 

12.  F  =  (2xz  -  y2  +  yzexvi:)i  -  (2xy  +  xzexyz)\ 

+  (x2  +  xyexyz)  k 

13.  F  =  (2x  +  y)i  +  (zcosyz  +  x)j  +  (ycosyz)k 

14.  F  =  (y  +  z)i  +  2zj  +  (x  +  y)k 

15.  F  =  ex  s'my  i  +  e*  cos  y  j  +  (3z2  +  2)  k 

16.  F  =  3x2i+ —  j +  2zlnyk 

y 

17.  F  =  (e-yz  -  yzexyz)i  +  xz(e-yz  +  exyz)i 

+  xy(e~yz  -  e"z)  k 

1 8.  Of  the  two  vector  fields 

F  =  xy2z3  i  +  2x2y  j  +  3x2y2z2  k 

and 

G  =  2xy  i  +  (x2  +  2yz)  j  +  y2  k 

one  is  conservative  and  one  is  not.  Determine  which 
is  which,  and  for  the  conservative  field  find  a  scalar 
potential  function. 

19.  (a)  Let  /  be  a  function  of  class  C1  defined  on  a  con- 

nected domain  in  R" .  Show  that  if  the  gradient  of 
/  vanishes  at  all  x  =  (xi, . . . ,  x„)  in  its  domain, 
then  /  is  constant. 


6.3  |  Exercises  449 


(b)  Suppose  that  F  is  a  conservative  vector  field  de- 
fined on  a  connected  subset  of  R" .  Show  that  if  g 
and  h  are  both  class  C1  potential  functions  for  F, 
then  g  and  h  must  differ  by  a  constant. 

20.  Find  all  functions  M(x,  y)  such  that  the  vector  field 

F  =  M(x ,  y )  i  +  (x  sin  y  —  y  cos  x)  j 
is  conservative. 

21.  Find  all  functions  N(x,  y  )  such  that  the  vector  field 

F  =  (ye2*  +  3xV)i+  y)j 
is  conservative. 

22.  Let  G(x,  y)  =  (xc*  +  y2)i  +  xy  j.  Find  all  functions 
g(jc)  such  that  the  vector  field  F  =  g(x)G  is  conserva- 
tive on  all  of  R2. 

23.  Find  all  functions  N(x,  y ,  z)  such  that  the  vector  field 

F  =  (x3y  —  3x2z)  i  +  N(x,  y ,  z)  j  +  (2yz  -  x3)  k 
is  conservative. 

24.  For  what  values  of  the  constants  a  and  b  will  the  vector 
field 

F  =  (3x2  +  3y2z  sinxz)i  +  (aycosxz  +  bz)\ 
+  (3xy2  sinxz  +  5y)k 
be  conservative? 

25.  Let  F  =  x2  i  +  cos  y  sin  z  j  +  sin  y  cos  z  k. 

(a)  Show  that  F  is  conservative  and  find  a  scalar  po- 
tential function  /  for  F. 

(b)  Evaluate  fxF-ds  along  the  path  x:  [0,  1]  ->  R3, 
x(t)  =  (t2  +  \,et,e2t). 

Show  that  the  line  integrals  in  Exercises  26-28  are  path  inde- 
pendent, and  evaluate  them  along  the  given  oriented  curve  and 
also  by  means  of  Theorem  3.3. 


26. 


L 


(3x  —  5y)  dx  +  (7y  —  5x)  dy ;  C  is  the  line  segment 


from  (1,3)  to  (5,  2). 


2j     f  x  dx 

Jc  Jx 


x  dx  +  y  dy 
Jx^+ 


C  is  the  semicircular  arc  of  x2  + 


28 


=  4  from  (2,  0)  to  (-2,  0). 
.  j  (2y  —  3z)  dx  +  (2x  +  z)  dy  +  (y  —  3x)  dz\  C  is  the 

line  segment  from  the  point  (0,  0,  0)  to  (0,  1,  1)  fol- 
lowed by  the  line  segment  from  the  point  (0,  1,  1)  to 
(1,2,3). 

In  Exercises  29— 32,  find  the  work  done  by  the  given  vector  field 
F  in  moving  a  particle  from  the  point  A  to  the  point  B  whose 
coordinates  are  as  indicated. 


29.  F  =  (3x2y  -  y2)i  +  (x3  -  2xy)j;  A(0,  0),  B{2,  1) 

30.  F  =  3^yi  +  2x3/2j;A(l,2),  S(9,  1) 

31.  F  =  (2xyz-y2z3)i  +  (x2z-2xyz3)j 

+  (x2y  -  3xy2z2)k;  A(l,  1,  1),  5(6,  4,  2) 

32.  F  =  2xy  cos  z  i  +  x2  cos  z  j  —  x2y  sin  z  k; 

A(l,  l,7r/2),  S(2,  3,0) 

33.  (a)  Determine  where  the  vector  field 

X  +  xv2        x2  +  1 

F  =  — — 1  —t 

r  r 

is  conservative. 

(b)  Determine  a  scalar  potential  for  F. 

(c)  Find  the  work  done  by  F  in  moving  a  particle  along 
the  parabolic  curve  y  =  1  +  x  —  x2  from  (0,  1)  to 
(1,1)- 

34.  Let  /,  g,  and  h  be  functions  of  class  C1  of  a  single 
variable. 

(a)  Show  that  F  =  (/(x)  +  y  +  z)  i  +  (x  +  g(y)  + 
z)  j  +  (x  +  y  +  h(z))  k  is  conservative. 

(b)  Determine  a  scalar  potential  for  F.  (Your  answer 
will  involve  integrals  of  /,  g,  and  h.) 

(c)  Find  fc  F  •  ds,  where  C  is  any  path  from 
(x0,  y0,  zo)  to  (xi,  yi,  zi). 

35.  Consider  the  vector  field 

F  =  (2x  +  z)  cos  (x2  +  xz)  i  -  (z  +  1)  sin  (y  +  yz)  j 

+  (x  cos  (x2  +  xz)  —  y  sin  (y  +  yz))  k. 
(a)  Determine  if  F  is  conservative. 


(b)  If  x(f)  =  ^3,  t2.  nt  -  sin y  I,  0<r  <  1,  evalu- 
ate fxF-ds. 

36.  Consider  the  vector  field 

G  =  (2x  +  z)  cos  (x2  +  xz)  i 

+  (x-(z  +  l)sin(y  +  yz))j 

+  (x  cos  (x2  +  xz)  -  y  sin  (y  +  yz))  k. 

(a)  How  is  G  different  from  the  vector  field  F  in  Ex- 
ercise 35?  Is  G  conservative? 

(b)  If  x(f)  =  ^r3,  r2, 7rf  -  sin  ^J,  0  <f  <  1,  evalu- 
ate fxG-ds. 

37.  Let  F  be  the  gravitational  force  field  of  a  mass  M  on  a 
particle  of  mass  m: 


7r  r 


GMm 


(x2  +  y2  +  z2)3/2 


(xi  +  yj  +  zk). 


450       Chapter  6  |  Line  Integrals 


(This  is  the  force  field  of  Example  3  in  §3.3.)  Given  xo  =  (xq,  yo,  zo)  to  Xi  =  (jci,  yi,  zi)  depends  only  on 

that  G,  M,  and  m  are  all  constants,  show  that  the  ||xo||  and  ||xi||. 

work  done  by  F  as  a  particle  of  mass  m  moves  from 


True/False  Exercises  for  Chapter  6 


1 .  If  C  is  the  parabola  y  =  4  —  x2  with  —  2  <  x  <  2,  then 
Jc  y  sinx  ds  =  0. 

2.  If  F  =  — i  +  j  +  k  and  C  is  the  straight  line  from  the 
origin  to  (2,  2,  2),  then  f„  F  •  ds  =  2«/3, 

3.  If  F  =  xi  +  y]  +  zk  and  C  is  the  straight  line  from 
(3,  3,  3)  to  the  origin,  then  fc  F  •  ds  is  positive. 

4.  Suppose  that  f(x)  >  0  for  all  x.  Let  F  =  fix) i.  If  C 
is  the  horizontal  line  segment  from  ( 1 ,  l)to(2,  l),then 
fcF-d$>  0. 

5.  Suppose  that  f(x)  >  0  for  all  x.  Let  F  =  f(x)  i.  If  C 
is  the  vertical  line  segment  from  (0,  0)  to  (0,  3),  then 
fc¥-ds>  0. 

6.  If  x  is  a  unit-speed  path,  then  f  F  •  ds  =  f  (F  •  v)  ds, 
where  v  denotes  the  velocity  of  the  path. 

7.  If  x  and  y  are  two  one-one  parametrizations  of  the 
same  curve  and  F  is  a  continuous  vector  field,  then 
fxY.ds  =  fyF.ds. 

8.  If  a  nonvanishing,  continuous  vector  field  F  is  every- 
where tangent  to  a  smooth  curve  C,  then  F  does  no 
work  along  the  curve. 

9.  If  a  nonvanishing,  continuous  vector  field  F  is  every- 
where normal  to  a  smooth  curve  C,  then  F  does  no 
work  along  the  curve. 

10.  If  the  curve  C  is  the  level  set  at  height  c  of  a  func- 
tion f(x,  y),  then  fc  f(x,  y)ds  is  c  times  the  length 
of  C. 

11.  If  f(x,y,z)  is  a  continuous  function  and 
Jc  f{x,  y,  z)ds  =  0  for  all  curves  C  in  R3,  then 
f(x,  y,  z)  =  0  for  all  (x,  y,  z)  G  R3. 

12.  If  a  closed  curve  C  is  a  level  set  of  a  function  f(x,  y) 
of  class  C1  and  V/  /  0,  then  the  flux  of  V/  across  C 
is  always  zero. 

13.  If  a  closed  curve  C  is  a  level  set  of  a  function  f(x,  y) 
of  class  C1  and  V/  /  0,  then  the  circulation  of  V/ 
along  C  is  always  zero. 

14.  If  a  vector  field  F  has  constant  magnitude  3  and  makes 
a  constant  angle  with  a  curve  C,  then  the  work  done 
by  F  along  C  is  3  times  the  length  of  C. 


15.  If  F  is  a  continuous  vector  field  everywhere  tangent  to 
an  oriented  C1  curve  C,  then  Jc  F  •  ds  =  fc  ||F||  ds. 

16.  IfF  is  a  constant  vector  field  on  R2,  then  <pc  F -ds  =  0, 
where  C  is  any  simple,  closed  curve. 

17.  IfF  is  an  incompressible  (i.e.,  divergenceless)  C  vec- 
tor field  on  R2  and  C  is  a  simple,  closed  curve,  then 
the  circulation  of  F  along  C  is  always  zero. 

18.  IfF  is  an  incompressible  C1  vector  field  on  R2,  then 
the  flux  across  any  simple,  closed  curve  C  in  R2  is 
always  zero. 

19.  If  C  is  a  simple  curve  in  R2,  then  f„  V/  •  ds  =  0. 

20.  If  C  is  a  simple,  closed  curve  in  R2  and  /  is  of  class 
C1,thenjfcV/-Js  =  0. 

21 .  F  =  (ex  cos  y  +  3)i  —  ex  sin  y  j  is  a  conservative  vec- 
tor field  on  R2. 

22.  If  /  and  g  are  functions  of  class  C1  defined  on  a  region 
D  in  R2,  then 

f   fVg-ds=£>  gVf-ds. 

J  3D  JdD 

23.  If  C  is  a  closed  curve  in  R3  such  that  <fc  F  •  ds  =  0, 
then  F  is  conservative. 

24.  (fc  x  dx  +  y  dy  +  z  dz  =  0  for  all  simple,  closed 
curves  C  in  R3 . 

25 .  <fc  ex  (cos  y  sin  z  dx  +  sin  y  sin  z  dy  +  cos  y  cos  z)  dz 
=  0  for  all  simple,  closed  curves  C  in  R3. 

26.  If  V  x  F  =  0,  then  F  is  conservative. 

27.  Let  M(x,  y)  and  N{x,  y)  be  C1  functions  with  domain 
R2  -  {(0,  0)}.  If  dM/dy  =  dN/dx,  then  §c  M  dx  + 
N  dy  =  0  for  any  closed  curve  C  in  R2. 

28.  Let  M(x,y,z),  N(x,y,z),  and  P(x,y,z)  be  C1 
functions  with  domain  R3  -  {(0,  0,  0)}.  If  dM/dy  = 
8N/dx,  dM/dz  =  dP/dx,  and  8N/dz  =  dP/dy, 
then  <fc  M  dx  +  N  dy  +  P  dz  =  0  for  any  closed 
curve  C  in  R3 . 

29.  If  F:R"  — >  R",  then  there  is  at  most  one  function 
/:  R     R  such  that  V/  =  F. 

30.  If  F  is  a  differentiable  vector  field  and  F  =  V  x  G, 
then  V  •  F  =  0. 


Miscellaneous  Exercises  for  Chapter  6  451 


Miscellaneous  Exercises  for  Chapter  6 


Let  C  C  R"  be  a  piecewise  Cl  curve  and  /:XCR"  — >  R 
a  continuous  function  whose  domain  X  includes  C.  Then  we 
define  the  quantity 


[/]. 


fcfds  fcfds 


,  ds       length  of  C 


to  be  the  average  value  of  f  along  C.  Exercises  1—5  concern 
the  notion  of  average  value  along  a  curve. 

1 .  Explain  why  it  makes  sense  to  use  the  preceding  inte- 
gral formula  to  represent  the  average  value.  (A  careful 
explanation  involves  the  use  of  Riemann  sums.) 

2.  Suppose  that  a  thin  wire  is  shaped  as  a  helical  curve 
parametrized  by 

x(t)  =  (cost,  sin?,  t),    0  <  t  <  3tt. 

If  fix,  y,  z)  =  x2  +  y2  +  2z2  +  1  represents  the  tem- 
perature at  points  along  the  wire,  find  the  average  tem- 
perature. 

3.  Find  the  average  y-coordinate  of  points  on  the  upper 
semicircle  y  =  y/a2  —  x2. 

4.  Find  the  average  z-coordinate  of  points  on  the  broken- 
line  curve  pictured  in  Figure  6.41. 


zS(x,  y,  z)ds. 


Using  these  definitions,  find  the  coordinates  of  the 
center  of  mass  of  the  wire. 

7.  Suppose  that  a  wire  is  bent  in  the  shape  of  a  quarter 
circle  of  radius  a.  Find  the  center  of  mass  of  the  wire 
if  the  density  at  points  on  the  wire  vary  as  the  square 
of  the  distance  from  the  center  of  the  wire. 

8.  (a)  Find  the  centroid  of  the  helical  wire  x  =  3  cost, 

y  =  3  sinf,  z  =  At,  where  0  <  t  <  4jt.  (Hint:  No 
calculation  should  be  necessary.) 

(b)  Find  the  center  of  mass  of  the  same  wire  if  the  den- 
sity at  each  point  of  the  wire  is  equal  to  the  square 
of  the  point's  distance  from  the  origin. 

If  a  thin  wire  is  bent  in  the  shape  of  a  curve  C  in  the  xy -plane 
and  has  mass  density  at  each  point  along  the  curve  given  by  a 
continuous  function  S(x,  y),  then  we  may  define  the  moments 
of  inertia  of  C  about  the  x-  and  y-axes,  respectively,  by 


f 


y  S(x,  y)  ds, 


Iy 


L 


x  S(x,  y)  ds, 


and  the  corresponding  radii  of  gyration  of  C  as 


'(2,1,0) 


Figure  6.41  The  broken-line  curve  of 
Exercise  4. 

5.  Find  the  average  value  of  f(x,  y,  z)  =  z2  +  xey  on  the 
curve  C  obtained  by  intersecting  the  (elliptic)  cylinder 
x2/5  +  y2  =  1  by  the  plane  z  =  2y. 

6.  A  metal  wire  is  bent  in  the  shape  of  the  semicircle 
x2  +  y2  =  4,  y  >  0,  lying  in  the  xy-plane.  Suppose 
that  the  mass  density  at  each  point  (x,  y,  z)  of  the  wire 
is  S(x,  y,  z)  =  3  —  y. 

(a)  Find  the  total  mass  of  the  wire. 

(b)  Using  formulas  analogous  to  those  in  §5.6,  we  de- 
fine the  (first)  moments  of  the  wire  to  be 


/  xS(x, 
Jc 


y,  z)ds, 


j  yS(x,  y,  z)ds, 


ry 


where  M  denotes  the  total  mass  M  =  fc  S(x,  y)  ds  of  the  wire. 
Additionally,  the  moment  of  inertia  of  C  about  the  origin  (or, 
equivalently,  about  the  z-axis,  if  we  think  of  the  xy -plane  as 
embedded  in  R3,)  is 


j  (x2  +  y2)S(x,y)ds, 


with  corresponding  radius  of  gyration  r-  =  *JIZ/M.  Exer- 
cises 9-13  concern  moments  of  inertia  of  curves. 

9.  (a)  Consider  the  wire  of  Exercise  6  again.  Find  its  mo- 
ment of  inertia  about  the  y-axis. 

(b)  What  is  the  radius  of  gyration  r-  for  the  wire  about 
the  z-axis? 

10.  Find  the  moment  of  inertia  Ix  and  the  radius  of  gy- 
ration rx  about  the  x-axis  of  a  straight  wire  between 
(—2,  1)  to  (2,  3)  whose  density  varies  along  the  wire 
as  8(x,  y)  =  y. 

1 1 .  Find  the  moment  of  inertia  lx  and  the  radius  of  gyra- 
tion rx  about  the  x-axis  of  a  wire  shaped  as  the  curve 
y  =  x2  between  (0,  0)  and  (2,  4)  and  whose  density 
varies  as  S(x,  y)  =  x. 

12.  (a)  Suppose  that  a  thin  metal  wire  is  bent  into  a 

curve  C  in  R3  and  has  mass  density  at  each  point 


452       Chapter  6  |  Line  Integrals 


(x,  y,  z)  along  the  curve  given  by  a  continuous 
function  S(x,  y,  z).  Give  general  formulas  analo- 
gous to  those  in  §5.6  for  the  moments  of  inertia  of 
the  wire  about  the  three  coordinate  axes. 

(b)  Find  the  moments  of  inertia  about  the  coordinate 
axes  of  a  homogeneous  (i.e.,  constant  density) 
wire  shaped  like  the  helix  x  =  3  cos  t,  y  =  3  sin  t , 
z  =  At,  where  0  <  t  <  Air.  What  are  the  radii  of 
gyration? 

1 3.  Find  the  moment  of  inertia  /-  about  the  z-axis  of  a  wire 
in  the  shape  of  the  line  segment  between  (—  1 ,  1,2) 
and  (2,  2,  3)  if  the  density  along  the  segment  varies  as 
S(x,  y,  z)  =  1  +  z2.  What  is  rzl 

14.  Let  r  =  f(6)  be  the  polar  equation  of  a  curve  in  the 
plane. 

(a)  Use  scalar  line  integrals  to  show  that  the  arclength 
of  the  curve  between  (/(a),  a)  and  (f(b),  b)  is 

V(/(0))2  +  (fWde. 


r 


(b)  Sketch  the  curve  r  =  sin  (8 /2)  and  find  its  length. 

15.  (a)  Give  a  formula  in  polar  coordinates  for  the  scalar 
line  integral  of  a  function  g(x,  y  )  along  the  curve 
r  =  f(8),  a  <8  <  b. 

(b)  Compute  fc  gds,  where  g(x,  y)  =  x2  +  y2  —  2x 
and  C  is  the  segment  of  the  spiral  r  =  eie,  0  < 
6  <  2ji. 

Let  C  be  a  piecewise  C  ,  simple  curve  in  R3.  The  total  curva- 
ture K  of  C  is 


K 


j  k  ds, 


where  k  denotes  the  curvature  of  C.  (See  §3.2  to  review  the  no- 
tion of  curvature.)  Exercises  16-20  involve  the  notion  of  total 
curvature. 

16.  Show  that  if  C  is  a  simple  curve  of  class  C1 
parametrized  by  x(r),  a  <  t  <  b,  then 


J  a 


v  x  a 


dt. 


(Recall  that  v  and  a  denote,  respectively,  the  velocity 
and  acceleration  of  the  path  x.) 

1 7.  Find  the  total  curvature  of  the  helix 

x(?)  =  (3  cost,  3  sinf,  At),  Q<t<\0n. 

18.  Find  the  total  curvature  of  the  parabola  y  =  Ax2, 
A  >  0,  for  a  <  x  <  b. 

19.  Fenchel's  theorem  states  that  if  C  is  a  simple,  closed 
C1  curve  in  R3,  then  K  >  2n  and,  moreover,  K  =  In 
if  and  only  if  C  is  a  plane  convex  curve.  (A  simple, 
closed  curve  C  is  convex  if  the  line  segment  joining 


any  two  points  of  C  lies  entirely  in  the  region  en- 
closed by  C.)  Verify  Fenchel's  theorem  for  the  ellipse 

x2/a2  +  y2/b2  =  1. 

20.  Let  C  be  a  simple,  closed  C1  curve  in  R3.  Suppose  that 
the  curvature  k  of  C  is  bounded  (i.e.,  0  <  k  <  I /a  for 
some  a  >  0). 

(a)  Show  that  if  L  denotes  the  length  of  C,  then 
L  >  2na. 

(b)  What  can  you  say  about  C  if  L  =  2nal 

21.  Calculate  the  work  done  by  the  vector  field  F  = 
sin  x  i  +  cos  yj+.xzkona  particle  moving  along  the 
path  x(r)  =  (t1,  -t2,  t),  where  0  <  t  <  1. 

22.  Use  Green's  theorem  to  find  the  work  done  by  the  vec- 
tor field  F  =  x2y  i  +  (x  +  y)y  j  in  moving  a  particle 
from  the  origin  along  the  y-axis  to  the  point  (0,  1), 
then  along  the  line  segment  from  (0,  1)  to  (1,  0),  and 
then  from  (1,0)  back  to  the  origin  along  the  x-axis. 
(Warning:  Be  careful.) 

23.  Use  Green's  theorem  to  recover  the  formula 


f 


for  the  area  A  of  the  region  D  described  in  polar  coor- 
dinates by 

D  =  {(r,  6)  |  0  <  r  <  f(6),  a  <6  <  b}. 

24.  Let  C  be  a  piecewise  C  ,  simple,  closed  curve  in  R2. 
Show  that 


f{x)dx  +  g(y)dy  =  0, 


where  /  and  g  are  any  single-variable  functions  of 
class  C1. 

25.  Let  D  be  a  region  in  R2  whose  boundary  dD  consists 
of  finitely  many  piecewise  Cl,  simple,  closed  curves 
oriented  so  that  D  is  on  the  left  as  you  travel  along  any 
segment  of  3D.  If  (x,  y)  denotes  the  coordinates  of  the 
centroid  of  D,  show  that 

x  =   ^   £)   x2  dy 

2  ■  area  of  D  J3D 

and 


1 


area  of  D  JdD 


xy  dy. 


Also  show  that 


and 


1 


area  of  D  JdD 
1 


xy  dx 


y2  dx. 


2  ■  area  of  D  J3D' 

26.  Use  the  results  of  Exercise  25  to  find  the  centroid  of  the 
triangular  region  with  vertices  (0,  0),  (1 ,  0),  and  (0,  2). 


Miscellaneous  Exercises  for  Chapter  6  453 


27.  Use  the  results  of  Exercise  25  to  find  the  centroid  of 
the  region  in  R2  that  lies  inside  the  circle  of  radius  6 
centered  at  the  origin  and  outside  the  two  circles  of 
radius  1  centered  at  (4,  0)  and  (—2,  2),  respectively. 


28. 


Let  C  be  a  piecewise  C1,  simple,  closed  curve,  ori- 
ented counterclockwise,  enclosing  a  region  D  in  the 
plane.  Let  n  be  the  outward  unit  normal  vector  to  D. 
If  f(x,  y)  and  g(x,  y)  are  functions  of  class  C2  on  D, 
establish  Green's  first  identity: 


(/V2g  +  V/-Vg)dA 


fVg-nds. 


29.  Under  the  hypotheses  of  Exercise  28,  prove  Green's 
second  identity: 


(fV2g-gV2f)dA 


(fVg-gVf)-nds. 


A  function  g(x,  y)  is  said  to  be  harmonic  at  a  point  (xq,  yo)  if 
g  is  of  class  C2  and  satisfies  Laplace 's  equation 

2  &h 
dxL 


+ 


dy2 


0 


on  some  neighborhood  of(xo ,  yo).  We  say  that  g  is  harmonic  on 
a  closed  region  OCR2  if  it  is  harmonic  at  all  interior  points 
ofD  (i.e.,  not  necessarily  along  dD).  Exercises  30- 33  concern 
some  elementary  results  about  harmonic  functions  in  R2. 

30.  Suppose  that  D  is  compact  (i.e.,  closed  and  bounded) 
and  that  3D  is  piecewise  C1  and  oriented  as  in  Green's 
theorem.  Let  n  denote  the  outward  unit  normal  vector 
to  3D  and  let  dg/dn  denote  Vg  •  n.  (The  term  dg/dn 
is  called  the  normal  derivative  of  g.)  Use  Green's  first 
identity  (see  Exercise  28)  with  f(x,  y)  =  1  to  show 
that,  if  g  is  harmonic  on  D,  then 


9g 
dn 


ds  =  0. 


31. 


32. 


Let  /  be  harmonic  on  a  region  D  that  satisfies  the 
assumptions  of  Exercise  30.  Show  that 

,.9/ 


V/-V/JA 


dn 


ds. 


Suppose  that  f(x,  y)  =  0  for  all  (x,  y)  e  3D.  Use  Ex- 
ercise 31  to  show  that  f(x,  y)  =  0  throughout  all  of 
D.  (Hint:  Consider  the  sign  of  V/  •  V/.) 

33.  Let  D  be  a  region  that  satisfies  the  assumptions  of  Ex- 
ercise 30.  Use  the  result  of  Exercise  31  to  show  that  if 
fx  and  fi  are  harmonic  on  D  and  f\(x,  y)  =  fi(x,  y) 
on  3D,  then,  in  fact,  f  =  fa  on  all  of  D.  Thus,  we  see 
that  harmonic  functions  are  determined  by  their  values 
on  the  boundary  of  a  region.  (Hint:  Consider  f\  —  fi  ) 

34.  We  call  a  vector  field  F  on  R3  radially  symmetric  if 

it  can  be  written  in  spherical  coordinates  in  the  form 
F  =  f(p)ep,  where  ep  is  the  unit  vector  that  points  in 
the  direction  of  increasing  p-coordinate.  (See  §1.7.) 


(a)  Give  an  example  of  a  (nontrivial)  radially  sym- 
metric vector  field,  written  in  both  Cartesian  and 
spherical  coordinates. 

(b)  Show  that  if  /  is  of  class  C1  for  all  p  >  0,  then 
the  radially  symmetric  vector  field  F  =  f(p)ep  is 
conservative. 

-yi  +  xj 


35.  LetF 


x2  +  y2 

(a)  Verify  Green's  theorem  over  the  annular  region 
D  =  {(x,y)|a2<Jc2-|-y2<  1}. 
(See  Figure  6.42.) 


Figure  6.42  The  annular  region 
of  Exercise  35(a). 

(b)  Now  let  D  be  the  unit  disk.  Does  the  formula  of 
Green's  theorem  hold  for  D?  Can  you  explain  why? 

(c)  Suppose  C  is  any  simple,  closed  curve  lying  out- 
side the  circle 


Ca  =  {(x,y)\x2  +  y2  =  a2} 


Figure  6.43  The  curves  C  and  C„  of 
Exercise  35(c). 

(See  Figure  6.43 . )  Argue  that  if  C  is  oriented  coun- 
terclockwise, then 


x  dy  —  y  dx 
x2  +  y2 


2jt. 


454       Chapter  6  I  Line  Integrals 


36.  Consider  the  vector  field  F  =  — r         i  H — -   j. 

x2  +  y2       x2  +  y2 

(a)  Calculate  V  x  F. 

(b)  Evaluate  (fc  F  •  ds,  where  C  is  the  unit  circle 
x2  +  y2  =  1. 

(c)  Is  F  conservative? 

(d)  How  can  you  reconcile  parts  (a)  and  (b)  with 
Theorem  3.5? 

37.  (a)  Let  F  =  ey  i  +  x4  j.  Calculate  the  flux  fc  F  •  nds 

of  F  across  the  boundary  of  the  rectangle  R  = 
[0,  1]  x  [0,  5]. 

(b)  Let  /  and  g  be  of  class  C1  and  let  F  =  f(y)i  + 
g(x )  j .  Show  that  the  flux  of  F  across  any  piecewise 
C1  simple,  closed  curve  is  zero. 

38.  Use  Newton's  second  law  of  motion  F  =  ma.  to  show 
that  the  work  done  by  a  force  field  F  in  moving  a  par- 
ticle of  mass  in  along  a  path  x(f )  from  x(a)  to  x(b)  is 


equal  to  the  change  in  kinetic  energy  of  the  particle.  In 
other  words, 

j^F-d$  =  \m{v(b))2  -  \m(v(a))2, 

where  v(t)  =  \\v(t)\\ ,  the  speed  at  time  t.  (Use  the  prod- 
uct rule  for  dot  products  of  vector- valued  functions.) 

39.  Let  F  be  a  conservative  vector  field  on  R3  with  F  = 
— VV.  If  a  particle  travels  along  a  path  x,  recall  that 
its  potential  energy  at  time  t  is  defined  to  be  V(x(f)). 
Use  line  integrals  to  prove  the  law  of  conservation  of 
energy:  As  a  particle  of  mass  m  moves  between  any 
two  points  A  and  B  in  a  conservative  force  field,  the 
sum  of  the  potential  and  kinetic  energies  of  the  particle 
remains  constant.  (Use  Exercise  38  and  Theorem  3.3.) 
The  use  of  line  integrals  provides  an  alternative  proof 
of  Theorem  4.2  in  §4.4. 


Surface  Integrals 
and  Vector  Analysis 


7.1  Parametrized  Surfaces 

7.2  Surface  Integrals 

7.3  Stokes's  and  Gauss's 
Theorems 

7.4  Further  Vector  Analysis; 
Maxwell's  Equations 

True/False  Exercises  for 
Chapter  7 

Miscellaneous  Exercises  for 
Chapter  7 


7.1   Parametrized  Surfaces 

Introduction   

Surfaces  in  R3  may  be  presented  analytically  in  different  ways.  Here  are  two 
familiar  descriptions: 

1.  as  a  graph  of  a  function  of  two  variables,  that  is,  as  points  (x,  y,  z)  in  R3 
satisfying  z  =  f(x,  y)  (e.g.,  z  =  x2  +  Ay2); 

2.  as  a  level  set  of  a  function  of  three  variables,  that  is,  as  points  (x,  y,  z) 
such  that  F(x,  y,z)  =  c  for  some  suitable  function  F  and  constant  c  (e.g., 

z2y5 


7 

x  y 


-2-5  +  x  =  1). 


Both  of  these  descriptions  are  problematical.  As  noted  in  §2.1,  many  common 
surfaces  cannot  be  described  as  graphs  of  functions  of  two  variables.  Recall,  for 
example,  that  the  full  sphere  x2  +  y2  +  z2  =  1  is  not  the  graph  of  a  function  of 
two  variables.  Therefore,  description  1  is  not  sufficiently  general. 

There  are  also  problems  with  description  2.  Not  all  equations  of  the  form 
F(x,  y,z)  =  c  have  solutions  that  fill  out  surfaces.  Indeed  although  the  level  set 
of  F(x,  y,  z)  =  x2  +  y2  +  z2  at  height  1  is  a  sphere,  at  height  0  it  is  a  single  point, 
and  at  height  —  1  completely  empty.  In  addition,  it  is  somewhat  tricky  to  describe 
surfaces  (i.e.,  two-dimensional  objects)  in  R"  by  using  level  sets  when  n  is  larger 
than  3.  Another  approach  is  desirable  for  presenting  surfaces  analytically,  in  order 
to  avoid  the  problems  just  mentioned,  to  emphasize  clearly  the  two-dimensional 
nature  of  a  surface  and  to  facilitate  subsequent  calculations.  With  this  discussion 
in  mind,  we  state  the  following  definition: 


DEFINITION  1 .1  Let  D  be  a  region  in  R2  that  consists  of  a  connected  open 
set,  possibly  together  with  some  or  all  of  its  boundary  points.  A  parametrized 
surface  in  R3  is  a  continuous  function  X:flcR2-^R3  that  is  one-one  on 
D,  except  possibly  along  3D.  We  refer  to  the  image  X(D)  as  the  underlying 
surface  of  X  (or  the  surface  parametrized  by  X)  and  denote  it  by  S.  (See 
Figure  7.1.) 


The  restrictions  on  the  region  D  and  the  map  X  of  Definition  1.1  are  meant  to 
ensure  that  D  is  a  two-dimensional  subset  of  R2  with  a  two-dimensional  image. 


456       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


c 

X(D)  =  S 


Figure  7.1  A  parametrized  surface. 

If  we  write  the  component  functions  of  X,  then,  for  (s,  ?)  e  D, 

X(s,  t)  =  (x(s,  ?),  y(s,  ?),  z(s,  ?)), 
and  the  underlying  surface  S  can  be  described  by  the  parametric  equations 

x  =  x(s,  ?) 

y  =  y(s,  ?)     (s,  t)  e  D. 
z  =  z(s,  t) 


(1) 


EXAMPLE  1    Consider  the  parametrized  surface  X:  R2 
X(*,0  =  .(i-j)  +  f(i  +  2k)  +  3j. 


R  described  by 


The  image  of  X,  shown  in  Figure  7.2,  is  the  plane  through  the  point  (0,  3,  0), 
determined  by  the  vectors  a  =  i  —  j  and  b  =  i  +  2k.  (See  Proposition  5.1 
of  §1.5.)  ♦ 


(0,3,0)  Y 
=  X(0,  01.  v 


Figure  7.2  The  parametrized  plane  of  Example  1. 

EXAMPLE  2   Let  D  =  [0,  2tt)  x  [0,  n]  and  consider  X:  D  ->  R3  given  by 

X(s,  t)  =  (a  coss  sin?,  a  sins  sin/,  a  cos?). 
The  corresponding  parametric  equations  are 
x  =  a  cos  s  sin  / 

y  =  a  sins  sin?      0  <  s  <  2n,  0  <  ?  <  n. 
z  =  a  cos  ? 

The  parametric  equations  imply  that  x2  +  y2  +  z2  =  a2,  meaning  all  the  points 
of  S  =  X(D)  lie  on  a  sphere  of  radius  a  centered  at  the  origin.  The  paramet- 
ric equations  are  precisely  the  spherical-rectangular  coordinate  conversions  (see 
§1.7)  with  the  p -coordinate  held  constant  at  a  and  with  s  and  ?  used  instead  of  6 
and  (p.  Hence,  the  image  of  X  is  indeed  all  of  the  sphere.  (See  Figure  7.3.)  ♦ 


7.1  j  Parametrized  Surfaces  457 


t  z 

Image  of 


Figure  7.3  A  sphere  rendered  as  a  parametrized  surface. 


EXAMPLE  3   The  points  of  the  surface  parametrized  by 

x  =  a  cos  s 
■  y  =  a  sin  s      0  <  s  <  2jt 
z  =  t 

satisfy  the  equation  x2  +  y2  =  a2  and  so  can  be  seen  to  form  an  infinite  cylinder 
of  radius  a.  Figure  7.4  shows  how  the  function  X(s,  /)  =  (a  coss,  a  sins,  t)  maps 
the  infinite  strip  D  =  {(s,  t)  |  0  <  s  <  2n]  onto  the  cylinder  by  "gluing"  the  line 
s  =  0  to  s  =  2ir .  ♦ 


Figure  7.4  The  map  X  glues  together  the  edges  of  D  to  form 
a  cylinder. 


EXAMPLE  4  Let  D  c  R2  be  an  open  set,  possibly  together  with  some  or  all  of 
its  boundary  points.  If  /:  D  ->  R  is  a  continuous  scalar- valued  function  of  two 
variables,  then  it  is  not  difficult  to  parametrize  the  graph  of  /:  We  let 

R3 


X:  D  ->  R3  be 
That  is,  the  parametric  equations 


X(s,  t)  =  (s,  t,  f(s,  0). 


0,  0  e  D 

describe  the  points  of  the  graph  of  /.  (See  Figure  7.5.) 


x  =  s 
y  =  t 

z  =  f(s,  t) 


Coordinate  Curves,  Normal  Vectors, 

and  Tangent  Planes   

Let  S  be  a  surface  parametrized  by  X:  D  —>  R3.  If  we  fix  /  =  to  and  let  only  s 
vary,  we  obtain  a  continuous  map 

s  i — >  X(j,  to), 


458       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


D 


X(£>)  is  graph  of  z  =f(x,  y) 


Figure  7.5  The  graph  of  z  =  f(x,  y)  as  a  parametrized  surface. 

whose  image  is  a  curve  lying  in  S.  We  call  this  curve  the  s-coordinate  curve  at 
t  =  t0.  Similarly,  we  may  fixj  =  i0  and  obtain  a  map 

t  i — >  X(s0,  t), 

whose  image  is  the  f -coordinate  curve  at  s  =  so-  Figure  7.6  suggests  the  appear- 
ance of  the  coordinate  curves. 


s  =  s0 

 7  f 

D 

t  =  tn 


X(M„) 


X(s„,0 


Figure  7.6  The  coordinate  curves  of  a  parametrized  surface. 
EXAMPLE  5   The  points  of  the  parametrized  surface  T  defined  by 

X  =  (a  +  &  COS  0  COS  J  n^.^o 

;    ,  ,       :  .  0  <  s  <  2tt,  0  <  t  <  2jt: 

y  =  (a  +  ocosr)sms  ~               _  .  ~  , 

,   .  a,  b  positive  constants  with  a  >  b, 

z  =  b  sin  t  F 

satisfy  the  equation 

(y/x2  +  y2-ay  +  z2=b2. 
The  ^-coordinate  curve  at  t  =  0  is 


x  =  (a  +  b)  cos  s 
V  =  (a  +  b)  sin  s 
z  =  0 


and  is  readily  seen  to  be  the  circle  of  radius  a  +  b,  centered  at  the  origin  and  lying 
in  the  xy-plane.  In  general,  you  may  check  that  the  s-coordinate  curve  at  t  =  t0 
is  a  circle  of  radius  a  +  b  cos  to  (which  varies  between  a  —  b  and  a  +  b)  in  the 
horizontal  plane  z  =  b  sinfo-  (See  Figure  7.7.) 


7.1  |  Parametrized  Surfaces  459 


s-coordinate 
curve  at  t  =  kI2 


s-coordinate 
curve  at  t  =  0 


s-coordinate 
curve  at  t  =  n 


Figure  7.7  Some  s-coordinate  curves  for  the  torus  T  of  Example  5. 


The  f-coordinate  curve  at  s  =  0  is 


x  =  a  +  b  cos  f 
y  =  0 
z  =  bsmt 


The  image  is  a  circle  of  radius  b  centered  at  (a,  0,  0)  in  the  xz-plane  (i.e.,  the 
plane  y  =  0).  At  s  =  so,  the  f-coordinate  curve  is 


x  =  cos  so  {a  +  b  cos  t) 
y  =  sin  so  (a  +  b  cos  t)  . 
z  =  bsint 


Along  this  curve,  we  have  y/x  =  tans0,  a  constant.  The  curve  lies  in  the  vertical 
plane  (sin  so)x  —  (cos  so)y  =  0.  Moreover,  it  is  not  hard  to  see  that  the  distance 
from  any  point  on  this  curve  to  the  point  P(a  cos  so,  a  sin  so,  0)  is  b,  and  the 
image  is  a  circle  of  radius  b  centered  at  P.  See  Figure  7.8  for  examples  of 
f-coordinate  curves. 


s  =  3nl2 


s  =  nl2 


kI2   k  3nl2  2n 

Figure  7.8  Some  f-coordinate  curves  for  the  torus  T  of  Example  5. 


The  aforementioned  surface  T  is  called  a  torus,  a  doughnut-shaped  surface 
shown  in  Figure  7.9.  It  is  generated  both  by  the  collection  of  the  s-coordinate 
curves  and  by  the  collection  of  the  f-coordinate  curves.  ♦ 

Suppose  that  X(s,  f)  =  (x(s,  f),  y(s,  f),  z(s,  f)),  where  (s,  f)  6  D,  is  a  dif- 
ferentiable  (or  C1)  map,  in  which  case  we  say  that  the  parametrized  surface 
S  =  X(D)  is  differentiable  (or  C1)  as  well.  Then  the  coordinate  curves  X(s,  fo) 
and  X(so,  f)  have  well-defined  tangent  vectors  at  points  (so,  fo)  in  D.  (See 
Figure  7.10.)  To  find  the  tangent  vector  Ts  to  the  s-coordinate  curve  X(s,  fo) 


460       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


5-coordinate 


/  curve  at  t  =  t0 


Figure  7.9  The  parametrized  torus  T. 


X(s0,t) 


Figure  7.10  The  tangent  vectors  Ts  and  Tf  to  the  coordinate  curves. 

at  (so,  to),  we  differentiate  the  component  functions  of  X  with  respect  to  s  and 
evaluate  at  (sq,  to): 

Ts(s0,  t0)  =  —(s0,  to)  =  t-(«o,  *o)  i  +  t-(«o,  to)]  +  —(so,  t0)  k.  (2) 
as  as  as  as 

Similarly,  the  tangent  vector  T,  to  the  r-coordinate  curve  X(so,  t)  at  (s0,  t0)  may 
be  calculated  by  differentiating  with  respect  to  /: 

3X  dx  dy  dz 

Tt(s0,  to)  =  —(so,  to)  =  —(so,  t0)i+  —(so,  to)}  +  —(so,  fo)k.  (3) 
at  at  at  at 


Since  Ts  and  T,  are  both  tangent  to  the  surface  S  at  (so,  to),  the  cross  product 
Ts(so,  to)  x  Tt(so,  to)  will  be  normal  to  S  at  (so,  to),  provided  it  is  nonzero. 


DEFINITION  1.2  The  parametrized  surface  S  =  X(D)  is  smooth  at 
X(so,  to)  if  the  map  X  is  of  class  C1  in  a  neighborhood  of  (so,  to)  and  if 
the  vector 

N(50,     =  T,(j0,  h)  x  T,(i0,  to)  +  0. 

If  5  is  smooth  at  every  point  X(so,  to)  €  S,  then  we  simply  refer  to  S  as  a 
smooth  parametrized  surface.  If  S  is  a  smooth  parametrized  surface,  we  call 
the  (nonzero)  vector  N  =  T,  x  T,  the  standard  normal  vector  arising  from 
the  parametrization  X. 


7.1  ;  Parametrized  Surfaces  461 


EXAMPLE  6  We  claim  that  the  torus  T  of  Example  5  is  smooth.  Recall  that  T 
is  given  as  the  image  of  the  map 

X:  [0,2tv]  x  [0,2*]  ->  R3, 

X(s,  /)  —  ((a  +  b  cos  / )  cos  s,  (a  +  b  cos  /)  sin  s ,  b  sin  /), 
where  a  >  b  >  0.  Then  from  formulas  (2)  and  (3),  we  have 

TsC?o>  fo)  =  —(fl  +  &cos/o)  sins0  i  +  (a  +    cos/o)cosso  j 

and 

T((so>  *o)  =  —b  sin/0  coss0  i  —  b  sin/0  sini0  j  +  bcosfo  k, 

so  that 

Ts  x  T,  =  (a  +  b  cos  fo)^  cos  to  cos  so  i  +  (a  +  ^  cos  ?o)^  cos  /o  sin  so  j 
+  (a  +  b  cos  fo)^  sin  t0  k 
=       +  b  cosfo)(cos?o  coss0  i  +  cosfo  sinso  j  +  sinf0  k). 

Since  a  >  b  >  0,  the  factor  b(a  +  b  cos  fo)  is  never  zero.  Furthermore,  since  the 
sine  and  cosine  functions  are  never  simultaneously  zero,  at  least  one  component 
of  Ts  x  T,  is  never  zero.  Hence,  the  torus  is  a  smooth  parametrized  surface.  ♦ 

s  =  —2  t  s  =  2  z 


t  =  n/4 


Figure  7.1 1  The  cone  z2  =  x2  +  y2  as  a  parametrized  surface. 

EXAMPLE  7  The  equation  z2  =  x2  +  y2  defines  a  cone  in  R3.  (See  Figure 
7.11.)  If  z  is  held  constant  (which  corresponds  to  slicing  the  surface  by  a 
horizontal  plane),  then  the  expression  x2  +  y2  is  constant.  Hence,  the  constant-z 
cross  sections  are  circles  of  radius  |z|  or  a  single  point  in  the  case  of  the  vertex. 
This  suggests  that  we  can  parametrize  the  cone  by  using  one  parameter  variable 
for  z  and  another  for  the  angle  around  the  z-axis.  Thus,  we  obtain  the  following 
equations: 

X  =  s  cos  t 
■  y  =  s  sin/    0<f<  2tt. 
z  =  s 

Then  we  have 


T, .  =  cos/  i  +  sinf  j  +  k    and    T,  =  —  s  sin/  i  +  s  cos  t  j. 


462       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Therefore, 


TsxT,= 


i  J 

cos  t  sin  t 
-ssint  scosf 


k 

1 

0 


=  —  s  cos  t  i  —  s  sin  t  j  +  s  k. 


Note  that  Ts  x  T,  =  0  when  (and  only  when)  s  =  0.  The  cone  fails  to  be  smooth 
just  at  its  vertex  (the  single  point  of  the  underlying  surface  corresponding  to 
s  =  0).  ♦ 

Examples  6  and  7  suggest  why  the  terminology  "smooth"  is  used:  Intuitively, 
a  parametrized  surface  is  smooth  at  a  point  if  it  has  no  sharp  "cusps"  or  "corners" 
there.  This  is  true  of  the  torus  but  not  of  the  cone,  which  has  a  singularity  at  its 
vertex. 

If  a  parametrized  surface  is  smooth  at  a  point  X(so,  to),  then  we  define  the 
tangent  plane  to  S  =  X(D)  at  the  point  X(so,  to)  to  be  the  plane  that  passes 
through  X(s0,  to)  and  has 

N($0,  to)  =  Ts(so,  to)  x  Tt(s0,  to) 

as  normal  vector.  To  write  an  equation  for  this  plane,  we  denote  (x,  y,  z)  by  the 
(variable)  vector  x.  Then  the  tangent  plane  equation  is 


N(s0,fo)-(x-X(so,fo))  =  0. 


(4) 


If  we  write  the  components  of  N(so,  to)  as  Ai  +  B  j  +  Ck  and  X(s0,  to)  as 
(x(s0l  t0),  y(s0,  to),  z(so,  to)),  then  we  may  expand  equation  (4)  to  obtain 

A(x  -  x(s0,  to))  +  B(y  -  y(s0,  t0))  +  C(z  -  z(s0,  t0))  =  0.  (5) 

EXAMPLE  8   Consider  once  again  the  parametrized  cone  of  Example  7  as 

X(s,  t)  =  (s  cost,  s  sinf,  s). 

From  the  calculations  in  Example  7,  the  cone  is  smooth  at  the  point  (0,  1,1)  = 
X(l,  n/2),  and  so  the  tangent  plane  exists  at  that  point.  We  have 


(l,f)=j  +  k    and  T,(l,f) 


so  that 


N 


=  T,(l,f)xTf(l,!)  = 


i 

0 
-1 


=  -j  +  k. 


Hence,  equation  (5)  can  be  applied  to  verify  that  an  equation  for  the  tangent 
plane  is 


or,  more  simply, 


0(x-0)-l(y-l)  +  l(z-l)  =  0 


z  =  y. 


EXAMPLE  9  If  f(x,  y)  is  of  class  C1  in  a  neighborhood  of  a  point  (x0,  y0) 
in  its  domain  D,  then  the  graph  of  /  is  a  smooth  parametrized  surface  at 
(xo,  yo,  f(xo,  yo))-  Recall  from  Example  4  that  the  graph  of  /  is  parametrized  by 


7.1     Parametrized  Surfaces  463 


X:  D  ->  R3, 


X(s,  t)  =  (s,  t,  f(s,  0).  Then 


Ts  =  i  H  k    and    T,  =  j  + 


3s 


so  that 


N  =  Ts  x  T,  =  -  —  i  -  —  j  +  k. 

ds  dt3 

Note  that  N  is  nonzero  at  any  point  (s,  t,  f(s,  t))  =  (x,  y,  fix,  y)). 
Next,  consider  the  surface  defined  as  the  level  set 


S  =  {(x,  y,  z)  |  F(x,  y,  z)  =  c,  c  constant}. 


z 


Figure  7.1 2  A  cube. 


If  F  is  of  class  C1  in  a  neighborhood  of  a  point  (x0,  Vo,  Zo)  £  S  and  VF(x0,  yo, 
z0)  ^  0,  then  the  implicit  function  theorem  (Theorem  6.5  of  §2.6)  implies  that,  in 
principle  at  least,  the  defining  equation  F(x,  y,  z)  =  c  of  S  always  can  be  solved 
locally  for  (at  least)  one  of  the  variables  x,  y,  or  z  in  terms  of  the  other  two. 
In  other  words,  under  the  given  assumptions  on  F,  the  level  set  S  is  locally  the 
graph  of  a  C1  function  of  two  variables.  It  is  important  to  remember  that  the 
idea  of  "solving  locally"  does  not  mean  that  all  points  of  S  can  be  described  as 
the  graph  of  a  single  function  (as  we  already  know  quite  well  in  the  case  of  the 
sphere*2  +  y2  +  z2  =  1,  for  example),  but  rather  that,  near  points  (xo,  yo,  Zo)  £  S 
where  V F(xq,  yo,  zo)  ^  0,  a  portion  of  S  may  be  described  as  a  graph.  Hence, 
graphs  of  C1  functions  of  two  variables  and  level  sets  of  C1  functions  of  three 
variables,  under  certain  smoothness  hypotheses,  are  locally  equivalent  descrip- 
tions for  surfaces.  Moreover,  since  the  graph  of  a  C1  function  is  a  smooth 
parametrized  surface,  we  may  shift  relatively  freely  among  our  three  descriptions 
for  surfaces.  ♦ 

Smooth  parametrized  surfaces  are  of  primary  importance  because  of  the  ease 
with  which  we  may  adapt  techniques  of  calculus  (particularly  integral  calculus) 
to  them.  But  we  are  also  interested  in  piecewise  smooth  parametrized  surfaces. 


DEFINITION  1 .3  A  piecewise  smooth  parametrized  surface  is  the  union  of 
images  of  finitely  many  parametrized  surfaces  X,  :  D;-  — >  R3,  i  =  1, . . . ,  m, 
where 

•  Each  Di  is  a  region  in  R2  consisting  of  a  connected  open  set,  possibly 
together  with  some  of  its  boundary  points  (for  the  most  part,  we  want  Dt 
to  be  an  elementary  region); 

•  Each  X,  is  of  class  C1  and  one-one  on  all  of  D,  ,  except  possibly  along 
9  A; 

•  Each  Sj  =  X,  (D,  )  is  smooth,  except  possibly  at  finitely  many  points. 


EXAMPLE  10  The  surface  of  a  cube  is  a  piecewise  smooth  parametrized 
surface.  It  is  the  union  of  its  six  faces,  each  one  of  which  is  a  smooth  parametrized 
surface,  namely,  a  portion  of  a  plane. 

More  explicitly,  suppose  that  a  cube's  faces  are  portions  of  the  planes  x  =  0, 
x  =  1,  y  =  0,y  =  l,z  =  0,andz  =  1  as  in  Figure  7. 12.  Then  we  may  parametrize 


464       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


the  cube's  faces  by  X,  :  [0,  1]  x  [0,  1]  ->  R3, =  1, . . . ,  6,  where 

Xl(s,t)  =  (0,s,t);       X2(s,t)  =  (\,s,t);       X3(s,t)  =  (s,0,t); 

X4(s,t)  =  (s,l,t);       X5(s,t)  =  (s,t,0);       X6(s,t)  =  (s,t,l). 

Each  map  X,  is  clearly  of  class  C1  and  one-one.  In  addition,  the  faces  have 
well-defined  nonzero  normal  vectors.  For  example,  for  both  Xi  and  X2, 

N1=N2  =  T5xT(=jxk  =  i. 

Similarly, 

N3  =  N4  =  i  x  k  =  -j    and   N5  =  N6  =  i  x  j  =  k. 

None  of  these  vectors  vanishes.  There  is  no  consistent  way  to  define  normal 
vectors  along  the  edges  of  the  cube  (where  two  faces  meet).  That  is  why  the  cube 
is  only  piecewise  smooth.  ♦ 

Area  of  a  Smooth  Parametrized  Surface  

Now,  we  use  the  notion  of  a  parametrized  surface  to  calculate  the  surface  area  of  a 
smooth  surface.  In  the  discussion  that  follows,  we  take  S  =  X(D)  to  be  a  smooth 
parametrized  surface,  where  D  is  the  union  of  finitely  many  elementary  regions 
in  R2  and  X:  D  ->  R3  is  of  class  C1  and  one-one  except  possibly  along  3D. 


x 


Figure  7.1 3  The  image  of  the  As  x  At  rectangle  in  D  is  approximately  a 
parallelogram  spanned  by  Ts(sq,  to) As  and  T,(sq,  to) At. 

The  key  geometric  observation  is  as  follows:  Consider  a  small  rectangular 
subset  of  D  whose  lower  left  corner  is  at  the  point  (s0,  t0)  e  D  and  whose  width 
and  height  are  As  and  Af ,  respectively.  The  image  of  this  rectangle  under  X  is  a 
piece  of  the  underlying  surface  S  that  is  approximately  the  parallelogram  with  a 
corner  at  X(so,  to)  and  spanned  by  the  vectors  Ts(so,  to) As  and  T,(so,  to) At.  (See 
Figure  7.13.)  The  area  A  A  of  this  piece  is 

AA  %  ||Ts(s0,  f0)As  x  Tf(i0,  t0)At\\  =  \\Ts(s0,  f0)  x  T,(s0,  t0)\\AsAt. 

Now,  suppose  D  =  [a,  b]  x  [c,  d];  that  is,  suppose  D  itself  is  a  rectangle. 
Partition  D  into  n2  subrectangles  via 

a  =  so  <  s\  <  ■  ■  ■  <  sn  =  b    and    c  =  to  <  t\  <  ■  ■  ■  <  t„  =  d. 

Let  Asi  =  Sj  —  and  Af,  =  tj  —  tj-\  for  i,  j  —  1, . . . ,  n.  Then  S  is  in  turn 
partitioned  into  pieces,  each  of  which  is  approximately  a  parallelogram,  assuming 
As,-  and  Atj  are  small  for  i,  j  =  1, . . . ,  n.  If  AAy  denotes  the  area  of  the  piece 


7.1     Parametrized  Surfaces  465 
of  S  that  is  the  image  of  the  ijth  subrectangle  of  D,  then 

n 

Surface  area  of  S  =  ^  AA,; 

n 

«  J]  ||T,(*,_i,*;_i)xT,(j/_1,fy_i)||Aj/A^. 

Therefore,  it  makes  sense  to  define 

Surface  area  of  S  =      lim      >    AAy  =  /    /    ||TsxT, ||  rfsrfr 

and,  in  general,  where  D  is  an  arbitrary  region  (i.e.,  not  necessarily  a  rectangle), 


Surface  area  of  S  =  I  I  ||TS  x  T,  ||     Jr.  (6) 


Formula  (6)  can  be  extended  readily  to  the  case  where  5  is  a  piecewise  smooth 
parametrized  surface  by  breaking  up  the  integral  in  an  appropriate  manner. 

EXAMPLE  1 1  We  use  formula  (6)  to  calculate  the  surface  area  of  a  sphere  of 
radius  a.  Recall  from  Example  2  that  the  map 

X(s,  t)  =  (a  coss  sinf,  a  sins  sinf,  a  cos  t),     0  <  s  <  2it,     0  <  t  <  it 

parametrizes  the  sphere  in  a  one-one  fashion.  Then 

Ts  =  —a  sin  s  sin  t  i  +  a  cos  s  sin  t  j 

and 

T,  =  a  cos  s  cos  t  i  +  a  sin  s  cos  f  j  —  a  sin  f  k. 

Hence, 


so  that 


Tj  xT,  =  — a2  sin  f  (cos  s  sin  f  i  +  sin s  sin  f  j  +  cos  /  k), 


\TS  x  T, ||  =  a2  sin?. 


Therefore,  formula  (6)  becomes 

r  f2n  ?         r  ? 

Surface  area  =  J     J     a  sin/ ds  dt  —  I    2na  sintdt 
Jo  Jo  Jo 

=  2ttci2(—  cos  t)\"  =  2na2(l  +  1)  =  Ana2. 

This  result  checks  with  the  well-known  formula  for  the  surface  area  of  a  sphere. 
Note,  however,  that  if  we  let  0  <  s  <  4jt,  0  <  t  <  ir,  then  the  image  of  X  is  the 
same  sphere,  but  formula  (6)  would  yield 

-Ait 

a2  sin?  ds  dt  =  in  a2. 


P1Z 

Jo  Jo 


This  "overcount"  is  why  we  must  assume  that  the  map  X  is  (nearly)  one-one. 
(Note:  The  aforementioned  parametrization  fails  to  be  smooth  at  t  =  0  and  at 
t  =  jt,  but  this  is  along  dD  and  so  does  not  affect  the  surface  area  integral.)  ♦ 


466       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


If  we  write  the  component  functions  of  X  as 

X(s,  f)  =  (x(s,  t),  y(s,  t),  z(s,  0), 

we  find  that 


Ts  x  T,  = 


i 

j 

k 

dx 

Oy 

dz 

97 

3  s 

3s 

dx 

dy 

dz 

It 

dt 

97 

dy  dz  dy  dz\  .  (  dx  dz  dx  dz 
ds  ~dl  ~  ~dtds)  1+  \  dlYs  ~  JsYt 

Using  the  notation  of  the  Jacobian,  we  obtain 

d(x,z) 


J  + 


dx  dy 
97  97 


dx  dy 
97  97 


NO,  t)  =  Ts  x  T, 


,  d(x,y) 
J  +  —  rk. 


(7) 


90,0      90,0"  '  30,0 

This  alternative  formula  for  the  normal  vector  to  a  smooth  parametrized  surface 
will  prove  useful  to  us  on  occasion.  For  the  moment,  we  take  its  magnitude: 


UNO,  Oil  = 


+ 


9(x,  z) 


90,  0  /      V  30,  0 
Hence,  formula  (6)  may  also  be  written  as 


+ 


9(y,  z) 
30, 0 


EXAM  PLE  1 2  Find  the  surface  area  of  the  torus  described  in  Example  5 .  Recall 
that  the  torus  is  parametrized  as 

x  =  (a  +  b  cos  0  cos  s 

y  =  (a  +  bcost)sins      0<s,t<2n,    a  >  b  >  0. 
z  =  b  sin  t 


Thus, 


3Q,  y) 
90,  0 


— (a  +  Z?  cos  0  sins 
(a  +b  cos  f )  cos  ^ 


-bsint  coss 
sin?  sin s 


(a  +  b  cos  f  )(£>  sin  /  sin2  s  +  bsint  cos2  s) 


d(x,  z) 
30,  0 


(a  +  Z?  cos  O^7  sint, 

-(a  +  Z?  cos  0  sins 
0 


-bsint  cos 5 
Z?cosr 


=  —(a  +  b  cos  0&  cos  t  sins, 


7.1  |  Exercises  467 


and 


3(y,  z) 


(a  +  &cosf)coss 
0 


3(5,  f) 

=  (fl  +  Z?  cos      cos  f  COS  5 
By  formula  (8),  we  have 
Surface  area 


bsint  sin s 
b  cos  r 


JO  JO 


(a  +  b  cos  r)2[Z?2  sin2  f  +  b2  cos2  /  sin2  s  +  b2  cos2  f  cos2  .?]  ds  dt. 


Using  the  trigonometric  identity  cos2  8  +  sin2  6  =  1  twice,  we  simplify  the  inte- 
gral to 

p  2jr    p  2.71  p  2tc 

I      I     (a  +  b  cos  t)b  ds  dt  =  I     2nb(a  +  b  cos  t)dt 
Jo    Jo  Jo 

=  2jib{at  +  b  sinf)^" 

=  4n2ab.  ^ 

EXAMPLE  1 3  Suppose  that  a  smooth  surface  is  described  as  the  graph  of  a  C1 
function /(x,  y),  that  is,  by  the  equation  z  =  f(x,  y),  where  (x,  y)  varies  through  a 
plane  region/).  Then  the  standard  parametrization  X(.5\  /)  =  (s,  t,  f(s,  r))implies 

T.s  x  T,  =  -fsi-  ft  j  +  k. 

(See  Example  9.)  Formula  (6)  yields 

Surface  area  =  j  j  \\TS  x  T(||  ds  dt  =  j j  J  f2  +  / 2  +  1  ds  dt. 

Since  x  =  s,  y  =  t  in  this  parametrization  of  the  graph,  we  conclude  that 


Surface  area  of  the  graph  of  f(x,  y)  over  D 


-IL 


f2  +  f2  +  ldxdy. 


(9) 


One  final  note:  It  is  not  at  all  clear  that  either  formula  (6)  or  formula  (8) 
depends  only  on  the  underlying  surface  S  =  X(D)  and  not  on  the  particular 
parametrization  X.  These  formulas  are  independent  of  the  parametrization,  as  we 
shall  observe  in  the  following  section,  in  the  context  of  general  surface  integrals. 


7.1  Exercises 


1 .  Let  X:  R2  —>  R3  be  the  parametrized  surface  given  by  (b)  Find  an  equation  for  the  plane  tangent  to  this  surface 

,  9      ?            9     „  ^  at  the  point  (3,  1,  1). 

X(s,t)  =  (s2  -t2,s  +  t,s2  +  3f).  F      v  ' 

2.  Find  an  equation  for  the  plane  tangent  to  the  torus 

(a)  Determine  a  normal  vector  to  this  surface  at  the 

point  =  ^    ^  cos  t)  cos  s,  (5  +  2  cos  t)  sin  s,  2sinf) 

(3,  1,1)  =  X(2,  -1).  at  the  point  ((5  -  V3)/V2.  (5  -  V3)/V2,  l). 


468       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


3.  Find  an  equation  for  the  plane  tangent  to  the  surface 

X  =  es,        y  =  t2e2s,        z  =  2e~s  + 1 
at  the  point  (1,4,0). 

4.  Let  X(s,  t)  =  (s2  cost,  s2  sinf,  s),  -3  <  s  <  3,  0  < 
t  <  In. 

(a)  Find  a  normal  vector  at  (s,  t )  =  (—  1 ,  0). 

(b)  Determine  the  tangent  plane  at  the  point  (1,  0,  —1). 

(c)  Find  an  equation  for  the  image  of  X  in  the  form 

F(x,y,z)  =  Q. 

5.  Consider  the  parametrized  surface  X(s,  f)  =  (s,  s2  + 
t,t2). 

(a)  Graph  this  surface  for  —  2  <  s  <  2,  —2  <  t  <  2. 
(Using  a  computer  may  help.) 

(b)  Is  the  surface  smooth? 

(c)  Find  an  equation  for  the  tangent  plane  at  the  point 
(1,0,1). 

6.  Describe  the  parametrized  surface  of  Exercise  1  by  an 
equation  of  the  form  z  =  f(x,y). 

7.  Let  S  be  the  surface  parametrized  by  x  =  scost, 
y  =  s  sinf,  z  =  s2,  where  s  >  0,  0  <  f  <  2n . 

(a)  At  what  points  is  S  smooth?  Find  an  equation  for 
the  tangent  plane  at  the  point  ( 1 ,  V3,  4). 

(b)  Sketch  the  graph  of  S.  Can  you  recognize  S  as  a 
familiar  surface? 

(c)  Describe  S  by  an  equation  of  the  form  z  =  f(x,  y). 

(d)  Using  your  answer  in  part  (c),  discuss  whether  S 
has  a  tangent  plane  at  every  point. 

8.  Verify  that  the  image  of  the  parametrized  surface 

X(s,  t)  =  (2  sin s  cos  f ,  3  sins  sinf,  coss), 
0  <  s  <  TV,     0  <  f  <  2jt, 

is  an  ellipsoid. 

9.  Verify  that,  for  the  torus  of  Example  5 ,  the  s  -coordinate 
curve,  when  f  =  to,  is  a  circle  of  radius  a  +  b  cos  to. 

1 0.  The  surface  in  R3  parametrized  by 

X(r,  6)  =  (rcos6,rsia6,  9),  r>0,  -oo<6<oo, 

is  called  a  helicoid. 

(a)  Describe  the  r-coordinate  curve  when  0  =  jr/3. 
Give  a  general  description  of  the  r -coordinate 
curves. 

(b)  Describe  the  ^-coordinate  curve  when  r  =  1. 
Give  a  general  description  of  the  8 -coordinate 
curves. 

(c)  Sketch  the  graph  of  the  helicoid  (perhaps  using 
a  computer)  for  0  <  r  <  1 ,  0  <  9  <  An .  Can  you 
see  why  the  surface  is  called  a  helicoid? 


11.  Given  the  sphere  of  radius  2  centered  at  (2,  —1,  0), 
find  an  equation  for  the  plane  tangent  to  it  at  the  point 
(1,  0,  V2)  in  three  ways: 

(a)  by  considering  the  sphere  as  the  graph  of  the 
function 

fix,  y)  =  j4-{x-2)2-{y+\)2; 

(b)  by  considering  the  sphere  as  a  level  surface  of  the 
function 

F(x,  y,  z)  =  (x-  2)2  +  (y  +  l)2  +  z2; 

(c)  by  considering  the  sphere  as  the  surface  para- 
metrized by 

X(s,  t )  =  (2  sin  s  cos  f  +  2,  2  sin  s  sin  f  —  1,2  cos  s ). 

In  Exercises  12—15,  represent  the  given  surface  as  a  piecewise 
smooth  parametrized  surface. 

12.  The  lower  hemisphere  x2  +  y2  +  z2  =  9,  including  the 
equatorial  circle. 

13.  The  part  of  the  cylinder  x2  +  z2  =  4  lying  between 
y  =  —  1  and  y  =  3. 

14.  The  closed  triangular  region  in  R3  with  vertices 
(2,0,  0),  (0,  1,0),  and  (0,0,  5). 

15.  The  hyperboloid  z2  —  x2  —  y2  =  1.  (Hint:  Use  two 
maps  to  parametrize  the  surface.) 

16.  This  problem  concerns  the  parametrized  surface 
X(s,t)  =  (s3,t3,st). 

(a)  Find  an  equation  of  the  plane  tangent  to  this  surface 
at  the  point  (1,  -1,  -1). 

(b)  Is  this  surface  smooth?  Why  or  why  not? 

(c)  Use  a  computer  to  graph  this  surface  for  —  1  <  s  < 

1,  -1  <  t  <  1. 

(d)  Verify  that  this  surface  may  also  be  described  by 
the  xyz-coordinate  equation  z  =  ffxy.  Try  using 
a  computer  to  graph  the  surface  when  described  in 
this  form.  Many  software  systems  will  have  trou- 
ble, or  will  provide  an  incomplete  graph,  which  is 
one  reason  why  parametric  descriptions  of  surfaces 
are  desirable. 

17.  The  surface  given  parametrically  by  X(s,  t )  = 
(st,t,s2)  is  known  as  the  Whitney  umbrella. 

(a)  Verify  that  this  surface  may  also  be  described  by 
the  jcyz-coordinate  equation  y2z  =  x2. 

(b)  Is  X  smooth? 

(c)  Use  a  computer  to  graph  this  surface  for —2  <  s  < 

2,  -2  <  f  <  2. 

(d)  Some  points  (x,  y,  z)  of  the  surface  do  not  cor- 
respond to  a  single  parameter  point  (s,  f).  Which 
ones?  Explain  how  this  relates  to  the  graph. 

(e)  Give  an  equation  of  the  plane  tangent  to  this  sur- 
face at  the  point  (2,  1,  4). 


7.2  |  Surface  Integrals  469 


(f)  Show  that  at  the  point  (0,  0,  1)  on  the  image  of  X 
it's  reasonable  to  conclude  that  there  are  two  tan- 
gent planes.  Give  equations  for  them. 

18.  Let  S  be  the  surface  defined  as  the  graph  of  a  function 
f(x,  y)  of  class  C1.  Then  Example  4  shows  that  S  is 
also  a  parametrized  surface.  Show  that  formula  (5)  for 
the  tangent  plane  to  S  at  (a ,  b ,  f(a ,  b))  agrees  with  that 
of  formula  (4)  in  §2.3. 

1 9.  (a)  Write  a  formula  for  the  tangent  plane  to  a  surface 

described  by  the  equation  y  =  g(x,  z). 

(b)  Repeat  part  (a)  for  a  surface  described  by  the  equa- 
tion x  =  h(y,  z). 

20.  Suppose  X:  D  — >  R3  is  a  parametrized  surface  that  is 
smooth  at  X(sq,  to).  Show  how  the  definition  of  the 
derivative  DX(sq,  to)  (see  Definition  3.8  of  Chapter  2) 
can  be  used  to  give  vector  parametric  equations  for  the 
plane  tangent  to  S  =  X(D)  at  the  point  X(sq,  to). 

21 .  Use  the  result  of  Exercise  20  to  provide  parametric 
equations  for  the  plane  tangent  to  the  surface  X(s,  t  )  = 
(s,  s2  +  t,  t2)  at  the  point  (1,0,  1).  Verify  that  your  an- 
swer is  consistent  with  that  of  Exercise  5(c). 

22.  Use  the  parametrization  in  Example  3  to  verify  that  the 
surface  area  of  a  cylinder  of  radius  a  and  height  h  is 
2nah. 

23.  Let  D  denote  the  unit  disk  in  the  jf-plane.  Let  X:  D  — > 
R3  be  defined  by  (s  +  t,  s  —  t ,  s ).  Find  the  surface  area 
ofX(D). 

24.  Find  the  surface  area  of  the  helicoid 

X:D^R3,    X(r,  6)  =  (r  cos0,r  sin 6,  6) 

for  0  <  r  <  1,  0  <  9  <  2nn,  where  n  is  a  positive 
integer. 

25.  A  cylindrical  hole  of  radius  b  is  bored  through  a  ball 
of  radius  a  (>  b  )  to  form  a  ring.  Find  the  outer  surface 
area  of  the  ring. 

26.  Find  the  area  of  the  portion  of  the  paraboloid  z  = 
9  —  x2  —  y2  that  lies  over  the  xy-plane. 

27.  Find  the  area  of  the  surface  cut  from  the  paraboloid 
z  =  2x2  +  2y2  by  the  planes  z  =  2  and  z  =  8. 


28.  Calculate  the  surface  area  of  the  portion  of  the  plane 
x  +  y  +  z  =  a  cut  out  by  the  cylinder  x2  +  y2  =  a2  in 
two  ways: 

(a)  by  using  formula  (6); 

(b)  by  using  formula  (9). 

29.  Let  S  be  the  surface  defined  by  the  equation  z  = 
f(x,  y).  If  f2  +  f2  =  a,  where  a  is  a  positive  con- 
stant, determine  the  surface  area  of  the  portion  of  S 
that  lies  over  a  region  D  in  the  .ty-plane  in  terms  of  the 
area  of  D. 

30.  Let  S  be  the  surface  defined  by 

1 

z  =    ,  for  z  >  1 . 

(a)  Sketch  the  graph  of  this  surface. 

(b)  Show  that  the  volume  of  the  region  bounded  by  S 
and  the  plane  z  =  1  is  finite.  (You  will  need  to  use 
an  improper  integral.) 

(c)  Show  that  the  surface  area  of  S  is  infinite. 

31 .  Find  the  surface  area  of  the  intersection  of  the  cylinders 

x2  +  y2  =  a2  and  y2  +  z2  =  a2. 

32.  Suppose  that  a  surface  is  given  in  cylindrical  coordi- 
nates by  the  equation  z  =  f(r,  0),  where  (r,  6)  varies 
through  a  region  D  in  the  r#-plane  where  r  is  non- 
negative.  Show  that  the  surface  area  of  the  surface  is 
given  by 

33.  Suppose  that  a  surface  is  given  in  spherical  coordi- 
nates by  the  equation  p  =  f(ip,  6),  where  (</s,  9)  varies 
through  a  region  D  in  the  ^6*-plane  and  f(<p,  9)  is  non- 
negative.  Show  that  the  surface  area  of  the  surface  is 
given  by 

x  7(/(?,  9)2  +  Mcp,  9)2)  sin2  <p  +  f0(<p,  9)2  d<p  d9. 


7.2   Surface  Integrals 

In  this  section,  we  will  learn  how  to  integrate  both  scalar-valued  functions  and 
vector  fields  along  surfaces  in  R3.  We  proceed  in  a  manner  that  is  largely  analo- 
gous to  our  explorations  of  line  integrals  in  §6.1:  We  begin  by  defining  suitable 
integrals  over  parametrized  surfaces  and  then  establish  that  the  particular  choice  of 
parametrization  doesn't  much  matter — that,  really,  only  the  underlying  surface  is 
important,  and  possibly  the  orientation. 


470       Chapter  7  I  Surface  Integrals  and  Vector  Analysis 


Area  AAt 


Figure  7.1 4  A  small  piece 
of  the  surface  S  has  area 
A  At.  The  point     is  located 
in  this  surface  piece. 


Scalar  Surface  Integrals   

Suppose  5  is  a  bounded  surface  in  R3  and  f(x ,  y,  z)  is  a  continuous,  scalar- valued 
function  whose  domain  includes  S.  Then  we  want  the  surface  integral  ffs  f  dS  to 
be  a  limit  of  some  kind  of  Riemann  sum.  So  suppose  S  is  partitioned  into  finitely 
many  small  pieces  and  that  the  area  of  the  kth  piece  is  AA*.  Let  c*  denote  an 
arbitrary  "test  point"  in  the  kth  piece.  (See  Figure  7. 14.)  Then  the  surface  integral 
of  /  over  S  should  be 


fffds=  „1t  nE/(c*)AA* 


(1) 


provided  of  course,  that  this  limit  exists. 

Now,  we  add  some  formalism  to  provide  a  proper  definition.  Suppose  that 
S  is  a  smooth  parametrized  surface;  that  is,  suppose  that  S  is  the  image  of  the 
C1  map  X:  D  — >•  R3,  where  D  is  a  connected,  bounded  region  in  R2.  Let  /  be  a 
continuous  function  defined  on  S  =  X(D).  As  seen  in  §7.1,  the  small  rectangle 
in  D,  having  dimensions  As  and  At  with  lower  left  corner  at  the  point  (sq,  to),  is 
mapped  by  X  to  a  piece  of  surface  that  is  approximately  a  parallelogram  of  area 

AA  sa  ||T,Oo,  to)  x  T,0o,  t0)\\AsAt. 

(See  Figure  7.13,  page  464.)  Suppose  D  is  the  rectangle  [a,  b]  x  [c,  d]  and  that 
we  partition  D  by 

a  =  sq  <  S\  <  ■  ■  ■  <  s„  =  b    and    c  =  to  <  t\  <  ■  ■  ■  <  tn  =  d. 

Then  the  limiting  sum  that  formula  (1)  represents  is 


lim       T  f(X(s*,  t*))  ||T,0,_i,        x  T/(s/_i,  Asj  Atj,  (2) 

all  Ajj.Afi-^O  r—1, 
i,j=l 

where 

Asi  =  Si  -  Si-i,     Atj  =  tj  -  tj-i,     Si-i  <  s*  <  Si,     and    t}-\  <  t*  <  tj. 

(Thus,  X(**,  tj)  is  an  arbitrary  point  in  the  image  of  the  subrectangle  [si-i,  st]  x 
[tj-i,  tj]  and,  hence,  a  "test  point"  in  the  corresponding  small  surface  piece.)  But 
then  the  limit  in  formula  (2)  is 

f    f  f(X(s,t))\\JsxJt\\dsdt. 

J  c     J  a 

When  D  is  a  more  arbitrary  region  than  a  rectangle,  it  makes  sense  to  use  the 
following  definition  for  the  surface  integral  of  a  function  over  a  parametrized 
surface: 


DEFINITION  2.1  Let  X:  D  — >  R3  be  a  smooth  parametrized  surface,  where 
D  c  R2  is  a  bounded  region.  Let  /  be  a  continuous  function  whose  domain 
includes  S  =  X(D).  Then  the  scalar  surface  integral  of  /  along  X,  denoted 
JfxfdS,  is 

j 'f  fdS  =  ff  f(X(s,  t))  \\TS  x  T,  ||  ds  dt 


-IL 


f(X(s,t))  \\N(s,t)\\dsdt. 


7.2  |  Surface  Integrals  471 


Although  we  need  not  assume  that  the  map  X  is  one-one  on  D  in  order  to 
work  with  the  integral  in  Definition  2.1,  in  practice  we  usually  find  it  useful  to 
take  X  to  be  one-one,  except  perhaps  along  dD.  If  this  is  the  case,  and  if  /  is 
identically  1  on  all  of  X(£)),  then 

if  fdS=  if  ldS=  if  ||  Ts  x  Tf  \\dsdt  =  surface  area  of  X(D), 
J  Jx  J  Jx  J  Jd 

as  stated  by  formula  (6)  in  §7. 1 .  The  scalar  surface  integral  in  Definition  2. 1  is  thus 

a  generalization  of  the  integral  we  use  to  calculate  surface  area.  We  can  think  of 

ffxfdS  as  the  limit  of  a  "weighted  sum"  of  surface  area  pieces,  the  weightings 

given  by  /.  If  /  represents  mass  or  electrical  charge  density,  then  ffxfdS  yields 

the  total  mass  or  total  charge  on  X(D)  (assuming  X  is  one-one,  except  perhaps 

along  dD). 

For  computational  purposes,  recall  that  if  we  write  the  components  of  X  as 
X(s,  t)  =  (x(s,  t),  y(s,  t),  z(s,  t)), 

then 

™   t\    t   ,t      d(y>z).     d(x,z).  d(x,y) 

N(i,  t)  =  To  x  T,  =  i  ■  H  k. 

v    '  d(s,  t)       d(s,  t)  J     d(s,  t) 

We  obtain 


f(x(s,  t),  y(s,  t),  z(s,  0) 


d(s,  t)  J     V  90>  0  /     V  9(s,  t) 


EXAMPLE  1  We  evaluate  / fx  z3  dS,  where  X:  [0,  2tt]  x  [0,  n]  -+  R3  is  the 
parametrized  sphere  of  radius  a : 

X(s,  /)  =  (a  coss  sinf,  a  sins  sinf,  a  cos  t). 

Using  Definition  2.1  or  its  reformulation  in  formula  (3),  we  find  that 


m°-m  =  i\jM)  +{hJT))  +l«Z7)j 

=  J a4(sin2  t  cos2  t  +  sin2  s  sin4 1  +  cos2  s  sin4 1) 
=  a2\J sin2  /  cos2  t  +  sin4 1 


=  a2J  sin2  ?  (cos2  t  +  sin2  f) 


=  a2  sinf. 
(See  also  Example  1 1  in  §7.1.)  Hence, 

'jr  pin 


/>  />  ptx    p Lit  p jr 

/  /  z3  dS  =  I     I     (a  cost)3a2  sint  ds  dt  =  a5  /    2tt  cos3  t  sin  f 
=  27rfl5  Kcos4f)|o  =2na5  (_I-(-I))=0. 


472       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


S*.  z  =  15 


Sl:x2  +  y1  =  9 


/  S2:z  =  0 

x 

Figure  7.15  The  closed 
cylinder  of  radius  3  and 
height  15  of  Example  2. 


To  define  and  evaluate  scalar  surface  integrals  over  piecewise  smooth  para- 
metrized surfaces,  simply  calculate  the  surface  integral  over  each  smooth  piece 
and  add  the  results. 

EXAMPLE  2   Let  5  be  the  closed  cylinder  of  radius  3  with  axis  along  the  z-axis, 
top  face  at  z  =  15,  and  bottom  face  at  z  —  0,  as  shown  in  Figure  7.15.  Then  5 
is  a  piecewise  smooth  surface;  it  is  the  union  of  the  three  smooth  parametrized 
surfaces  Si,  52,  and  53  described  next.  We  calculate  ffs  z  dS. 
The  three  smooth  pieces  may  be  parametrized  as  follows: 


5i  (lateral  cylindrical  surface): 


52  (bottom  disk): 


3  cos  s 
3  sins 


x  = 

y  = 

z  =  0 


s  cos? 
s  sin? 


0<s<2n,    0  <  /  <  15, 


0  <  s  <  3,    0  <  t  <  2n, 


and 


53  (top  disk): 


x  =  s  cos/ 

v  =  ssin?  0  <  s  <  3,  0  <  t  <  lit. 
z=  15 


Using  Definition  2.1,  we  have 

p  p  pl5  p2tt 

I  I  zdS=  I     I     1 1|(— 3  sins  i  +  3  coss  j)  x  k||  ds  dt 
J  Js,  Jo  Jo 


10  JO 

■15  p2n 


Jo  Jo 


1 1| 3  sins  j  +  3  coss  i||  ds  dt 


/i  15    pin  pi 

=         /     Itdsdt  = 
Jo    Jo  Jo 


15 


biztdt  =  3jrr  L  =  675n 


Now,  / fs  z  dS  =  0,  since  z  vanishes  along  the  bottom  of  5.  For  53,  we  have 


/  /  z.dS=  [  [  l5dS=  15-  /  / 
J  Js,  J  Js,  J  Js, 


l5dS=  15-  /  /  IdS 

is,  J  Js, 

15  •  area  of  disk  =  15  •  (9n)  =  \35n. 


Therefore, 


f  f  zdS=  f  f  zdS+  [  [  zdS+  f  f 
J  J s  J  Jsi  J  Js?  J  Js, 


zdS 


'  S\  J  J  Si  J  J  s, 

=  6757T  +  0  +  1357T  =  8107T. 

If  a  surface  5  is  given  by  the  graph  of  z  =  g(x,  y),  where  g  is  of  class  C1  on 
some  region  D  in  R2,  then  5  is  parametrized  by  X(jf,  y)  =  (x,  y,  g(x,  y))  with 
(x,  y)  e  D.  (See  Example  4  of  §7.1.)  Then,  from  Example  13  in  §7.1, 


so  that 


N(x,y)=  -g.Ti-gvj  +  k, 


jjjdS=  j j  f(x,  y,  g(x,  y))Jgl  +  g2y+ldx  dy. 


(4) 


7.2  |  Surface  Integrals  473 


;  „ 


D 


Figure  7.16  The  graph 
of  4  —  x2  —  y2  over  the 
disk  D  of  radius  2. 


Figure  7.17  If  u  =  tan  1  4, 

then  sec  u  =  y/VJ, 


EXAMPLE  3  Suppose  S  is  the  graph  of  the  portion  of  the  paraboloid  z  = 
4  —  x2  —  y2,  where  (x,  y)  varies  throughout  the  disk 


D  =  {(x,y)  eR  \  x  +  y  <  4}. 

(See  Figure  7.16.)  Formula  (4)  makes  it  straightforward  although  rather  involved 
to  calculate 

jj  (4  -  z)dS,     where  X(x,  y)  =  (x,  y,  4  -  x2  -  y2). 
In  particular,  we  have 

jj(4-z)dS  =  jj  (4  -  (4  -  x2  -  v2))  74jc2  +  4y2  +  1  dx  dy 

=  jj  (x2  +  v2V4jc2  +  4v2  +  1  dx  dy. 

To  integrate,  we  switch  to  polar  coordinates;  that  is,  we  let  x  =  r  cos  9  and  y  = 
r  sin#,  where  0  <  r  <  2,  0  <  0  <  2n.  The  desired  integral  becomes 


n7.7T      nT.    nA      n  ATT 

/      /    r2^4r2  +\rdrd6  =  /     /     r^Ar2  +\d9dr 
Jo    Jo  Jo  Jo 

=  2n  /    rV4r2+  \dr, 
Jo 

by  Fubini's  theorem.  Now  let  2r  =  tanw;  that  is,  let  r  =  \  tanw  so  that  dr 


-2  P2iT 


|  sec2  udu.  The  previous  integral  transforms  into 


2;r 


I 


tan-1  4 


tan 


3  u  •  y/ tan2  u  +  1  •  -  sec2  u  du 


8  Jo 

=  ~\ 

8  Jo 

=  -/ 

8  Jo 


tan3  u  sec3  u  du 


tan2  u  sec2  u  •  (sec  u  tan  udu) 


(sec2  u  —  1)  sec2  m  sec  u  tanudu. 


Now,  let  w  =  sec  m  so  d  w  =  sec  w  tan  «  .  Hence,  when  u  =  0,w  =  1  and  when 
m  =  tan-1  4,  «;  =  -^17-  (See  Figure  7.17.)  Thus,  the  w-integral  becomes 


71 


Vl7 


(w  —  \)w  dw 


71  r^<  * 


w  )dw 


7t   (\     5        1  3 

-id  iy 

5  3 


Tt       M2  i—     17  r— 

— VT7  Vl7  , 

5  3        J     \5  3 


71. 


1  1 


391V17  +  1 
60 


474       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Alternatively,  we  could  calculate  the  integral  /„2  r3>/4r2  +  1  dr  using  integra- 
tion by  parts  with  u  =  r2  (so  du  =  2rdr)  and  dv  =  r^/Ar2  +  1  dr  (so  v  = 


_l_(4r2  +  ^3/2) 


i/ 


Figure  7.1 8  The 

helicoid  of 
Example  4. 


Vector  Surface  Integrals   

Now  we  develop  a  means  to  integrate  vector  fields  along  surfaces,  beginning  with 
a  definition. 


DEFINITION  2.2  LetX:  D  — ►  R3  be  a  smooth  parametrized  surface,  where 
D  is  a  bounded  region  in  the  plane,  and  let  F(x,  y,  z)  be  a  continuous  vector 
field  whose  domain  includes  S  =  X(£>).  Then  the  vector  surface  integral 
of  F  along  X,  denoted  ffx  F  •  dS,  is 

yy  f.js= yy  fcxcs.o^nc*, 

where  N(s,  ?)  =  Ts  x  T,. 


As  with  line  integrals,  you  are  cautioned  to  be  careful  about  notation  for  surface 
integrals.  In  the  vector  surface  integral  ffx  F  •  dS,  the  differential  term  should  be 
considered  to  be  a  vector  quantity,  whereas  in  the  scalar  surface  integral  ffxfdS, 
the  differential  term  is  a  scalar  quantity  (namely,  the  differential  of  surface  area). 

EXAMPLE  4  Let  F  =  x  i  +  y  j  +  (z  -  2y)  k.  We  evaluate  ffx  F  •  dS,  where  X 
is  the  helicoid 

X(s,  t)  =  (s  cost,  s  sinf,  t),     0  <  s  <  1,     0  <  f  <  27T. 

The  helicoid  is  shown  in  Figure  7.18. 
We  have 

K,    t,      d(y,z)  .     d(x,z)  .     9(x,  y) 

N(s,  t)  =  i  l  H  k 

v     y      3(s,  f)       9(5, 0  J     d(s,  t) 


sin  f     s  cos  f 
0  1 


=  sin  t  i  —  cos  tj  +  sk. 
Using  Definition  2.2,  we  obtain 

-2ir  pi 


cos  t   —s  sin  t 
0  1 


j  + 


cos?  -i  sin( 
sin  t     s  cos  t 


ff¥-dS=f     I  F(X(s,t))-N(s,t)dsdt 
J  Jx  Jo  Jo 

=  j         (s  cos  t  i  +  s  sin  t  j 
Jo  Jo 

+  (t  —  2s  sin  t )  k)  •  (sin  ?  i  —  cos  t  j  +  s  k)  ds  dt 

p  lit      pi  n  27T 

=         /  (st  -  2s2  smt)ds  dt  =  /      (\s2t  —  \s3  sin?)  |  ^dt 
Jo    Jo  Jo 


-L ( 


sin 


t)  dt  =  (\t2  +  \  cos/)!^  =  it2. 


7.2  |  Surface  Integrals  475 


EXAMPLE  5  Let  f(x,  y)  be  a  scalar-valued  function  of  class  C1  on  a  bounded 
domain  D  c  R2.  Suppose  S  is  the  surface  described  as  the  graph  of  z  =  f(x,  y); 
that  is,  S  =  X(D),  where  X(x,  y)  =  (x,  y,  f{x,  y)).  Then 

N(x,v)=-/,i-/vj  +  k, 

so  that  Definition  2.2  becomes 

J£  F  •  dS  =  JJ  F(x,  y,  f(x,  y))  •  (-/,  i  -  /,  j  +  k)  dx  dy.  (5) 
Formula  (5)  will  prove  to  be  quite  useful.  ♦ 

Further  Interpretations   

As  is  the  case  for  vector  and  scalar  line  integrals,  there  is  a  connection  between 
vector  and  scalar  surface  integrals.  Suppose  X:  D  ->  R3  is  a  smooth  parametrized 
surface  and  F  is  continuous  on  S  =  X(D).  Let  N(s,  t)  =  Ts  x  Tf  be  the  usual 
normal  vector  and  let 

N(j,  t) 
n(s,  n  =   . 

UNO,  on 

That  is,  n  is  the  unit  vector  pointing  in  the  same  direction  as  N.  In  particular, 

NO,  0  =  UNO,  Oil  Ms,  t). 
Using  Definition  2.2,  we  have  that  the  vector  surface  integral  is 

j  j  ¥-dS  =  j  j  ¥(X(s,t))-N(s,t)dsdt 

F(XO,  0)  •  (UNO,  OllnO,  0)  ds  dt 
F(XO,0)-n0,0l|NO,0ll^  dt 

i 

(F-n)dS.  (6) 

/x 

Since  n  is  a  unit  vector,  the  quantity  F  •  n  is  precisely  the  component  of  F  in  the 
direction  of  n.  In  other  words,  formula  (6)  says  that  the  vector  surface  integral  of  F 
along  X  is  the  scalar  surface  integral  of  the  component  of  F  normal  to  S  =  X(£)). 
It  is  the  surface  integral  analogue  of  formula  (3)  of  §6.1,  which  states  that  the 
vector  line  integral  of  F  along  a  path  x  is  the  scalar  line  integral  of  the  component 
of  F  tangent  to  the  image  curve.  To  summarize,  we  have  the  following  results: 


//.: 
//.: 


Line  integrals: 

.ds=  j  (F  •  T)ds. 

J  X 

(V) 

Surface  integrals: 

///' 

dS  =  jj  (F-n)dS. 

(8) 

476       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Figure  7.1 9  The  amount  of  fluid  transported 
across  a  small  piece  of  S  during  a  brief  time 
interval  At  may  be  approximated  by  the  volume 
of  a  parallelepiped. 


As  noted  in  §6. 1 ,  when  x  is  a  closed  path,  the  quantity  /x(F  •  T)  ds  in  equation 
(7)  is  called  the  circulation  of  F  along  x.  It  measures  the  tangential  flow  of  F 
along  the  path.  On  the  other  hand  the  quantity  f  /X(F  •  n)  dS  in  equation  (8)  is 
known  as  the  flux  of  F  across  5  =  X(D).  If  we  think  of  F  as  the  velocity  vector 
field  of  a  three-dimensional  fluid,  then  the  flux  may  be  thought  of  as  representing 
the  rate  of  fluid  transported  across  S  per  unit  time,  as  we  now  see.  (You  may 
wish  to  compare  the  following  discussion  with  the  one  in  §6.2  concerning  the 
two-dimensional  flux  across  a  curve.) 

To  avoid  notational  confusion,  we  use  u  and  u  to  denote  the  parameter  vari- 
ables for  X  and  t  as  the  variable  representing  time.  Consider  a  small  piece  of  S, 
having  area  AS,  and  the  amount  of  fluid  transported  across  it  during  a  brief  time 
interval  At.  This  amount  is  the  volume  determined  by  F  during  At.  Figure  7.19 
suggests  that  if  both  AS  and  At  are  sufficiently  small,  then  this  volume  can  be 
approximated  by  the  volume  of  an  appropriate  parallelepiped.  Therefore, 

Amount  of  fluid  transported  «  volume  of  parallelepiped 

=  (height)  (area  of  base) 

=  F(X(w0,  v0))At  ■  n(w0,  v0)AS,  (9) 

since  the  height  of  the  parallelepiped  is  the  normal  component  of  FAf .  We  obtain 
the  average  rate  of  transport  across  the  surface  piece  during  the  time  interval  At 
by  dividing  (9)  by  Af : 

Average  rate  of  transport  R»  F(X(m0,  vo))  •  n(w0,  vo)AS.  (10) 

Now,  break  up  the  entire  surface  S  =  X(Z))  into  many  such  small  pieces  and  sum 
the  corresponding  contributions  to  the  rate  of  transport  in  the  form  given  in  (10). 
If  we  let  all  the  pieces  shrink,  then,  in  the  limit  as  all  AS  ->  0,  we  have  that  the 
total  average  rate  AM/  At  of  fluid  transported  during  Af  is  approximately 

AM       f  f 

Finally,  let  At  —>  0  and  define  the  (instantaneous)  rate  of  fluid  transport  to  be 

dM       f  f 


7.2  |  Surface  Integrals  477 


Reparametrization  of  Surfaces   

As  seen  in  §6.1,  scalar  and  vector  line  integrals  over  curves  depend  on  the  geom- 
etry of  the  curve  (and  possibly  its  direction),  rather  than  on  the  particular  way  in 
which  the  curve  may  be  parametrized.  Much  the  same  is  true  for  surface  integrals. 
We  begin  with  a  definition,  analogous  to  Definition  1.3  of  Chapter  6. 


DEFINITION  2.3  Let  X:  D,cR2^  R3  and  Y:  D2cR2^  R3  be  para- 
metrized surfaces.  We  say  that  Y  is  a  reparametrization  of  X  if  there 
is  a  one-one  and  onto  function  H:  —>  D\  with  inverse  H  D\  — >•  Z>2 
such  that  Y(s,  t)  =  X(H(j,  t)),  that  is,  such  that  Y  =  X  o  H.  If  X  and  Y  are 
smooth  and  H  and  H  1  are  both  of  class  C1 ,  then  we  say  that  Y  is  a  smooth 
reparametrization  of  X. 


EXAMPLE  6   The  helicoid  parametrized  by 


X  =  s  cos  t 
y  =  s  sin  t 
z  =  t 


0  <  s  <  1,    0  <t  <2tt 


may  also  be  described  as 


x  =  -  cos2f 
2 


v  =  -  sin2f 
2 


0  <  s  <  2,    0  <  t  <  7t. 


z  =  2t 


The  first  description  corresponds  to  a  map  X:  [0,  1]  x  [0,  2tt]  — >  R3  and  the  sec- 
ond to  a  map  Y:  [0,  2]  x  [0,  n]  — >•  R3.  It  is  not  difficult  to  see  that  if  we  make  the 
change  of  variables  by  letting 


u  =  -    and    v  =  2t, 


then 


Y(s,  t)  =  X(w,  v). 

Equivalently,  we  can  define  a  function  H:  [0,  2]  x  [0,  tt] 
H(s,  t)  =  (s/2,  2t).  Then  H  is  one-one  and  onto  and  Y  = 
a  reparametrization  of  X. 


■>  [0,  1]  x  [0,2jr]with 
X  o  H.  Therefore,  Y  is 
♦ 


EXAMPLE  7  Suppose  X  is  a  smooth  parametrized  surface.  Let  Y(s,  t)  = 
X(w,  v),  where  u  =  t,  v  =  s.  That  is,  Y  =  X  o  H,  where  H(s,  t)  =  (t,  s).  Then 
Y  is  a  (smooth)  reparametrization  that  appears  to  accomplish  little.  However,  if 
we  let  Ny  denote  the  usual  normal  vector  Ts  x  T,  =  dY/ds  x  dY/dt,  then  we 
have 

3Y  _  aX  9Y  _  ax 

ds       dv  dt       du  ' 

so  that 

3Y     9Y     9X     9X        9X  9X 


478       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


The  parametrized  surface  Y  is  the  same  as  X,  except  that  the  standard  nor- 
mal vector  arising  from  Y  points  in  the  opposite  direction  to  the  one  arising 
from  X.  ♦ 


The  calculation  in  Example  7  generalizes  thus:  Suppose  X  is  a  smooth 
parametrized  surface  and  Y  is  a  smooth  reparametrization  of  X  via  H,  mean- 
ing that 

Y(s,  t)  =  X(u,  v)  =  X(H(s,  /))■ 

Since  H  is  assumed  to  be  of  class  C1,  we  can  show  from  the  chain  rule  that  the 
standard  normal  vectors  are  related  by  the  equation 

d(u,  v) 

ny(^0  =  -^-±Nx(u,v).  (11) 
d(s,t) 

(See  the  addendum  at  the  end  of  this  section  for  a  derivation  of  formula  (11).) 
Formula  (11)  shows  that  NY  is  a  scalar  multiple  of  Nx.  In  addition,  since  H  is 
invertible  and  both  H  and  H  1  are  of  class  C1,  it  follows  that  the  Jacobian  of  H 
is  either  always  positive  or  always  negative.  (To  see  this,  note  that  both  H  o  H  1 
and  H  1  o  H  are  the  identity  function.  Hence,  the  chain  rule  may  be  applied  to 
show  that  the  derivative  matrix  DH(s,  t)  is  invertible  for  each  (s ,  t);  therefore,  its 
determinant,  which  is  the  Jacobian  of  H,  must  be  nonzero.  Since  the  determinant 
is  a  continuous  function  of  the  entries  of  H,  it  thus  cannot  change  sign.)  Hence, 
the  standard  normal  NY  either  always  points  in  the  same  direction  as  Nx  or  else 
always  points  in  the  opposite  direction  (Figure  7.20).  Under  these  assumptions,  we 
say  that  both  H  and  Y  are  orientation-preserving  if  the  Jacobian  d(u,  v)/d(s,  t) 
is  positive,  orientation-reversing  if  d{u,  v)/d(s,  t)  is  negative. 


Figure  7.20  If  Y  is  an  orientation-reversing  reparametrization  of  X,  then  NY 
points  opposite  to  Nx- 

The  following  result,  a  close  analogue  of  Theorem  1 .4,  Chapter  6,  shows  that 
smooth  reparametrization  has  no  effect  on  the  value  of  a  scalar  line  integral. 


THEOREM  2.4  Let  X:  D\  — »  R  be  a  smooth  parametrized  surface  and  /  any 
continuous  function  whose  domain  includes  X(Dj ).  If  Y:  D2  — »•  R3  is  any  smooth 
reparametrization  of  X,  then 


fL/ds=fLfds- 


7.2  |  Surface  Integrals  479 


The  proof  of  Theorem  2.4  appears  in  the  addendum  to  this  section.  With  this 
result,  we  can  define  the  scalar  surface  integral  over  a  smooth  surface  S  by  taking 
a  smooth  parametrization  X:  D  —>  R3  with  S  =  X(D)  that  is  one-one,  except 
possibly  on  3D.  Then  we  define  the  scalar  surface  integral  of  /  on  S  by 


It  is  a  fact  (which  we  shall  not  prove)  that  any  two  smooth  parametrizations  of 
S  must  be  reparametrizations  of  each  other,  so  Theorem  2.4  tells  us  that  any 
particular  choice  of  parametrization  we  might  make  does  not  matter.  We  need  to 
assume  that  X  is  (nearly)  one-one  to  ensure  that  the  integral  is  taken  only  once 
over  the  underlying  surface  S  =  X(D).  It  is  also  a  straightforward  matter  to  extend 
these  comments  to  give  a  definition  of  a  scalar  surface  integral  of  a  function  over 
a  piecewise  smooth  surface. 

Analogous  to  Theorem  1.5  of  Chapter  6,  the  following  result  (whose  proof 
is  in  the  addendum)  tells  us  that  smooth  reparametrizations  only  affect  vector 
surface  integrals  by  a  possible  sign  change. 


THEOREM  2.5  Let  X:  D\  —>  R3  be  a  smooth  parametrized  surface  and  F  any 
continuous  vector  field  whose  domain  includes  X(Dj).  If  Y:  D2  —>  R3  is  any 
smooth  reparametrization  of  X,  then  either 


if  Y  is  orientation-reversing. 


Because  of  Theorem  2.5,  it  is  a  more  subtle  and  involved  matter  to  define 
a  vector  surface  integral  over  a  smooth  surface  than  to  define  a  scalar  surface 
integral.  Given  a  smooth,  connected  surface,  we  need  to  choose  an  orientation 
for  it.  This  is  akin  to  orienting  a  curve  but,  perhaps  surprisingly,  is  not  always 
possible,  even  for  a  well-behaved  smooth  parametrized  surface,  as  Example  8 
illustrates. 

Here  is  a  formal  definition  of  orientability  of  a  smooth  surface. 


DEFINITION  2.6  A  smooth,  connected  surface  S  is  orientable  (or  two- 
sided)  if  it  is  possible  to  define  a  single  unit  normal  vector  at  each  point  of 
S  so  that  the  collection  of  these  normal  vectors  varies  continuously  over  S. 
(In  particular,  this  means  that  nearby  unit  normal  vectors  must  point  to  the 
same  side  of  S.)  Otherwise,  S  is  called  nonorientable  (or  one-sided). 


It  is  a  fact  (clearly  suggested  by  Figure  7.21)  that  a  smooth,  connected  ori- 
entable surface  S  has  exactly  two  orientations. 

If  S  happens  to  be  the  image  X(D)  of  a  smooth  parametrized  surface  X:  D  -> 
R3,  then  the  normal  vectors 


if  Y  is  orientation-preserving,  or 


Ni  =  Ts  x  T,    and    N2  =  T,  x  T.v  =  -Nj 


480       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Figure  7.21  The  connected  orientable  surface  S  shown 
with  its  two  possible  orientations. 


Figure  7.22  Some  r-coordinate 
curves  of  the  parametrized  Mobius 
strip. 


Figure  7.23  The  Mobius  strip  of 
Example  8. 


can  be  used  to  give  unit  vector  normal  vectors  ni  =  Ni/||Ni  ||  andii2  =  N2/HN2II 
that  point  in  opposite  directions.  It  is  tempting  to  think  that  ni  and  112  always 
provide  two  orientations  for  S.  However,  even  though  both  rij  and  n2  may  vary 
continuously  with  respect  to  the  parameters  s  and  t,  it  is  not  clear  that  they  must 
vary  continuously  and  consistently  with  respect  to  the  points  on  the  underlying 
surface  5.  Example  8  is  a  famous  instance  of  a  nonorientable  surface. 

EXAMPLE  8   The  surface  parametrized  by 


-( 


1  +  t  cos  -  J  COS  s 

2/ 

1  +  t  cos  sins 


0  <  s  <  2jt, 


4  <  t  <  i 
2  —  —  2 


z  =  t  sin 


is  called  a  Mobius  strip.  It  may  be  visualized  as  follows:  The  r-coordinate  curve 
at  s  =  sq  is 


)S  —  J  t  +  cos  SQ 


so" 
2 


^  t  +  sin  so 


k  <  t  <  \ 
2  —    —  2 


=0 

=(■  s 

=  (™f)< 

This  is  a  line  segment  through  the  point  (cos  s0 ,  sin  s0 ,  0)  and  parallel  to  the  vector 
s0  cos  (y)  i  +  sin s0  cos  (y)  j  +  sin  (y)  k. 


COS.' 


Several  such  coordinate  curves,  marked  with  the  direction  of  increasing  t,  are 
shown  in  Figure  7.22.  We  see  that  the  Mobius  strip  is  generated  by  a  moving  line 
segment  that  begins  (at  s  =  0)  lying  along  the  positive  x-axis,  rises  to  a  vertical 
position  with  center  at  (—1,  0,  0)whens  =  n,  and  then  falls  back  to  horizontal,  but 
with  direction  reversed  at  s  =  lit.  The  s -coordinate  curve  at  t  =  0  is  parametrized 
by 


X  =  coss 

y  =  sins        0  <  S  <  27T, 

z  =  0 


and  so  is  a  circle  in  the  xy-plane.  The  full  Mobius  strip  is  shown  in  Figure  7.23. 
You  can  make  a  physical  model  by  taking  a  strip  of  paper,  giving  it  a  half-twist, 
and  joining  the  short  ends. 


7.2  |  Surface  Integrals  481 


You  can  understand  the  gluing  process  analytically  by  noting  that  the  map 

X:[0,27r]x[-I,I]^R3 

defining  the  Mobius  strip  as  a  parametrized  surface  has  the  property  that  X(0,  t)  = 
X(2tt,  —t)  but  is  otherwise  one-one.  Therefore,  every  point  (0,  /)  on  the  left  edge 
of  the  domain  rectangle  [0,  2jt]  x  [—  |,  |]  is  mapped  to  the  point  (1  +  t,  0,  0)of 
the  Mobius  strip,  as  is  the  point  (2jt,  —t)  on  the  right  edge  of  the  rectangle.  (See 
Figure  7.24.) 


1/2 


-1/2 


In 


Figure  7.24  Gluing  the  ends  of  a  strip  of  paper  so  that  the  arrows  align 
provides  a  model  of  the  Mobius  strip. 

Now,  let's  investigate  the  orientability  of  the  Mobius  strip.  The  standard  nor- 
mal vector  is 


NO,  t)  =  Ts  x  T,  = 


We  have 


and 


d(y,z)  .     d(x,z)  .  d(x,y) 

 1  H  k 

90,  t) 


90,  t) 
.  s 
2 


-  (cos  s  +  2t  (cos 


90,  0 
2 


cos 


t)) 


+ 


1 


4  cos 


4  cos3  -  +  t  (1  + 


s  /  s 
cos  -  ( 1  + 1  cos  - 
2  V  2 


2 


coss 


cos2  s 


N(0,/)  =  -j 


(1  +  Ok, 


Figure  7.25  Traveling  once 
around  the  circular  path  on  the 
Mobius  strip  forces  the  normal 
vector  to  reverse  direction. 


N(2;r,  -0  =  -  -  j  +  (1  +  0  k  =  -N(0,  0- 

Therefore,  a  uniquely  determined  normal  vector  has  not  been  defined.  More 
vividly,  imagine  traveling  along  the  Mobius  strip  via  the  ^-coordinate  path  at 
t  =  0,  that  is,  along  the  circular  path 

x(j)  =  X(s,  0)  =  (coss,  sins,  0),     0  <  s  <  2n. 

Follow  the  standard  normal  N.  At  s  =  0,  it  is  N(0,  0)  =  — k,  but  by  the  time 
we  close  the  loop,  it  is  N(27T,  0)  =  k.  This  apparent  reversal  of  the  normal 
vector  means  that  the  strip  is  not  orientable  at  all — it  is  one-sided.  (See 
Figure  7.25.)  ♦ 

A  smooth,  orientable  surface  together  with  an  explicit  choice  of  orientation 
for  it  is  called  an  oriented  surface.  If  S  is  such  a  smooth  oriented  surface,  then 
we  define  the  vector  surface  integral  of  F  along  S  by  finding  a  suitable  smooth 


482       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


parametrization  X  of  S  such  that  the  unit  normal  vector  N(s,  f)/||N(s,  t)\\  arising 
from  the  parametrization  agrees  with  the  choice  of  orientation  normal.  We  take 
the  surface  integral  to  be 


F  •  dS. 


By  Theorem  2.5,  if  Y  is  any  orientation-preserving  reparametrization  of  X,  the 
value  of  //yF'  dS  is  the  same  as  ffx  F  •  dS,  so  this  notion  of  a  surface  inte- 
gral over  the  underlying  oriented  surface  S  is  well-defined.  Even  though  we  may 
perfectly  well  calculate  ffx  F  •  dS,  where  X  is  the  parametrized  Mobius  strip  of 
Example  8,  it  does  not  make  sense  to  consider  the  surface  integral  over  the  under- 
lying Mobius  strip,  since  there  is  no  way  to  orient  it.  Similarly,  the  interpretation 
of  the  vector  surface  integral  as  the  flux  of  F  across  the  surface  only  makes  sense 
once  an  orientation  of  the  surface  is  chosen.  Then  the  flux  measures  the  flow  rate, 
positive  or  negative,  depending  on  the  choice  of  orientation.  (See  Figure  7.26.) 


Figure  7.27  The  sphere 
x2  +  y2  +  z2  =  a2  oriented 
by  outward-pointing  unit 
normal  vectors. 


Figure  7.26  How  flux  depends  on  orientation.  On  the 
left,  the  surface  S  is  oriented  by  unit  normal  vectors  so 


that  F  •  n  is  positive  at  every  point.  Hence, 


,F-dS  = 


ffs  F  •  n  dS  is  positive.  On  the  right,  S  is  given  the 
opposite  orientation  so  that  the  flux  ffsF-dS<0. 

Another  reason  for  de-emphasizing  the  role  of  parametrization  in  surface 
integrals  is  that  we  can  often  exploit  the  geometry  of  the  underlying  surface  and 
vector  field  when  making  calculations.  If  S  is  a  smooth,  orientable  surface  and  n  a 
unit  normal  that  gives  an  orientation  of  S  (so,  in  particular,  n  is  understood  to  vary 
with  the  points  of  S),  then,  for  a  continuous  vector  field  F  defined  on  S,  we  have 


dS. 


If  we  can  determine  a  continuously  varying,  unit  normal  vector  at  each  point  of 
S  (for  example,  if  S  is  the  graph  of  a  function  f(x,  y)  of  two  variables  or  the 
graph  of  a  level  set  f(x,  y,  z)  =  c  of  a  function  of  three  variables),  then  there  is 
a  good  chance  that  the  surface  integral  can  be  evaluated  readily. 


EXAMPLE  9  Let  F  =  xi  +  yj  +  zkbea  radial  vector  field,  and  suppose  S  is 
the  sphere  of  radius  a  with  equation  x2  +  y2  +  z2  =  a2.  Orient  S  by  outward- 
pointing  unit  normal  vectors  as  shown  in  Figure  7.27.  We  calculate  the  flux  of 
F  across  S  in  two  ways:  (1)  by  means  of  a  parametrization  of  S  and  (2)  via 
geometric  considerations,  that  is,  without  resorting  to  an  explicit  parametrization 
of  the  sphere. 

For  approach  (1),  use  the  usual  parametrization  X  of  the  sphere: 


x  =  a  cos  s  sin  t 
y  =  a  sin  s  sin  t 
z  =  a  cos  t 


0  <  s  <  2tt,  0  <  t  <  it. 


7.2  |  Surface  Integrals  483 


The  standard  normal  vector  for  this  parametrization  is 

N(s,  ?)  =  —  a2  sin?  (coss  sin?  i  + sins  sin?  j  +  cos?  k). 

(This  normal  vector  is  calculated  in  Example  11  of  §7. 1 .)  If  we  normalize  N,  we 
find  that 

N(s,  0 

n(s,  t)  =   =  —(cos s  sin?  i  +  sin s  sin?  j  +  cos  ?  k). 

I|N(j,OII 

Thus,  n  is  inward-pointing  at  every  point  on  the  sphere.  Therefore,  we  must 
make  a  sign  change  when  we  evaluate  the  vector  surface  integral,  if  we  use  the 
parametrization  just  given.  Hence,  we  have 

f  f F-dS  =  -  f  f  F-dS  =  -  f    f  F(X(s,t))-~N(s,t)dsdt 
J  Js  J  Jx  Jo  Jo 


n 

Jo  Jo 


2it 

(a  cos  s  sin  ?  i  +  a  sin  s  sin  ?  j  +  a  cos  ?  k) 


•  (—a2  sin?  (cos  j  sin?  i  +  sins  sin?  j  +  cos?  k)  ds  dt 

pir  p2n 

"ill  0  0  0  0  0 

=  a    J     I     sin?  (cos  s  sin  ?  +  sin  s  sin  ?  +  cos  t)dsdt 
Jo  Jo 

pit      pT-TT  n7t 

=  cr"  /     /     sin?  ds  dt  =  lira3  I  smtdt=4jza3. 
Jo  Jo  Jo 

Now,  reconsider  this  calculation  along  the  lines  of  approach  (2).  Since  S  is  denned 
as  a  level  set  of  the  function  f(x,  y,  z)  =  x2  +  y2  +  z2,  normal  vectors  can  be 
obtained  from  the  gradient: 

V(x2  +  y2  +  z2)  =  2x  i  +  2y  j  +  2z  k. 

If  we  normalize  the  gradient,  then  we  have  unit  normal  vectors.  Thus, 

2x  i  +  2y  j  +  2z  k      2x  i  +  2y  j  +  2z  k     x  i  +  y  j  +  z  k 


y/Ax1  +  4y2  +  4z2  2v/x2~+y2~+ 


zr 


because  x2  +  y2  +  z2  =  a2  at  points  on  S.  (Note  that  n  is  always  outward- 
pointing.)  Therefore, 


JJ.'-M=JJ,r"a 


-II. 


s 

=  ff*2  +  y2  +  *\s=(fa-<ls 

J  Js         a  J  Js  a 

=  a  j  j  dS  =  a  ■  area  of  S  =  a(4na2)  =  Ancr" .  ^ 

All  of  the  preceding  remarks  concerning  scalar  and  vector  surface  integrals 
can  be  adapted  to  define  integrals  over  piecewise  smooth,  connected  surfaces. 
Simply  add  the  contributions  of  the  surface  integrals  over  the  various  smooth 
pieces.  The  only  issue  is  that  of  orientation,  but  assuming  that  each  of  the  smooth 


484       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


pieces  is  orientable,  then  it  is  possible  to  provide  an  orientation  to  the  surface 
as  a  whole.  Here's  how:  Suppose  Si  and  S2  are  two  smooth  surface  pieces  that 
meet  along  a  common  edge  curve  C.  Let  iij  and  n2  be  the  respective  unit  normal 
vectors  that  give  the  orientations  of  Si  and  S2  ■  Then  ni  and  n2  each  give  rise 
to  an  orientation  of  C  via  a  right-hand  rule.  (To  see  this,  for  j  =  1,  2,  point 
the  thumb  of  your  right  hand  along  n; ;  the  direction  of  your  fingers  will  indicate 
the  orientation  of  C.)  If  C  receives  opposite  orientations  from  nj  and  n2,  then  Si 
and  S2  are  oriented  consistently;  if  C  receives  the  same  orientation,  then  Si  and 
S2  are  oriented  inconsistently.  (See  Figure  7.28.) 


Figure  7.28  The  piecewise  smooth  surface  S  =  Si  U  S2 
oriented  consistently  on  the  left  and  inconsistently  on  the 
right. 


Figure  7.29  The 

piecewise  smooth 
cylindrical  surface  S 
of  Example  10  shown 
with  orientation 
normals. 


EXAMPLE  10  We  evaluate  / fs(x3  i  +  y3  j)  •  dS,  where  S  is  the  closed  cylinder 
bounded  laterally  by  x2  +  y2  =  4,  and  on  bottom  and  top  by  the  planes  z  =  0  and 
z  =  5,  oriented  by  outward  normal  vectors. 

Evidently,  S  is  the  union  of  three  smooth  oriented  pieces:  (1)  the  bottom 
surface  Si ,  which  is  a  portion  of  the  plane  z  =  0,  oriented  by  ni  =  — k;  (2)  the  top 
surface  S2,  which  is  a  portion  of  the  plane  z  =  5,  oriented  by  n2  =  k;  and  (3)  the 
lateral  cylindrical  surface  S3  given  by  the  equation  x2  +  y2  =  4  and  oriented  by 
normalizing  the  gradient  of  x2  +  y2  along  S3,  namely, 

2xi  +  2y\        xi  +  y  j       x'\  +  y\ 

113 


74jc2  +  4y2  t/x2^ 


r 


See  Figure  7.29  for  a  depiction  of  S. 
Now  we  calculate 


jjtfi  +  y'i).dS 


j jVi  +  y3j)-JS  + 


j ^(x3i+y3j).JS 


+  /  /  (*3i  +  y3j).JS 


i  is, 


+  y3j)-(-k)JS  +  JJ^xH  +  y3])-kdS 


+ 


lLixl+yi) 


dS 


0  +  0  + 


f  fsi 


\(x*  +  yA)dS. 


7.2  |  Surface  Integrals  485 

To  finish  the  evaluation,  we  may  parametrize  53  by 
x  =  2  cos  s 

y  =  2sins      0  <  s  <  2n,  0<t<5. 
vz  =  t 

Then 


jj(xH+y3j).dS=  f f  \(x*  +  yA)dS 


s3 

=  I    I     j(16cos4j  +  16  sin4    2  dt 
Jo  Jo 

=        /     16(cos4s  +  sin4  s)ds  dt 
Jo  Jo 

=  /    /     16((cos2s)2  +  (sin2s)2)ds^ 
Jo  Jo 

5    r2n       /  / j  +  cos2s\2  /l-COs2sx2N 


// 

Jo  Jo 


16   +   ds  dt, 


from  the  half-angle  substitution.  Thus, 

*5  sin 


f  f(xH  +  y3\)-dS=  f  f  f(2  +  2  cos2  2s)  ds  dt 
J  Js  Jo  Jo 


/>5  rln 

=  (8  +  4(1  +  cos  4s))  ds  dt. 

Jo  Jo 


By  once  again  using  the  half-angle  substitution,  we  get 

■5  pS 


f  f(xH  +  y3i)-dS=  f  (12i  +  sin4s)|2*  dt  =  f  24n  dt  =  120tt.  ♦ 
J  Js  Jo  Jo 

Summary:  Surface  Integral  Formulas 


Scalar  surface  integrals: 

For  a  surface  S  parametrized  by  X:  D  c  R2  R3, 

[  [  fdS  =  [  f  fdS=  f  f  f(X(s,t))\\TsxTt\\dsdt. 
J  Js  J  Jx  J  Jd 

Surface  area  element  is  dS  =  \\TS  x  Tr||  ds  dt. 

For  a  surface  S  described  as  a  graph  of  a  function  z  =  g(x,  y),  where  g:  D  c 
R2  R, 

ff  fdS  =  j j  fix,  y,  g(x,  y))  Jgx(x,  y)2  +  gy(x,  yf  +  1  dx  dy. 
Surface  area  element  is  dS  =  ^/ gx{x,  y)2  +  gy(x,  y)2  +  1  dx  dy. 


486       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Vector  surface  integrals: 

For  a  surface  S  parametrized  by  X:  D  C  R2  — >•  R3, 

jj^F-dS  =  jf(F-n)dS  =  j  j  F(X(s,  t)) -~N(s,  t)ds  dt, 

where  N  =  T,  x  T,  and  n  =  N/||N||. 

Vector  surface  integral  element  is  rfS  =  N(j,  t)ds  dt. 

For  a  surface  5  described  as  a  graph  of  a  function  z  =  g(x,  y),  where  g:  D  Q 


Rz  ->  R, 


Here  n  =  (-fo  i  -  gy  j  +  k)/^2  +  g2  +  1. 

Vector  surface  integral  element  is  S  =  (— gx  i  —  gy  j  +  k)dx  dy. 


Addendum:  Proofs  of  Theorems  2.4  and  2.5  — 

We  begin  by  establishing  formula  (1 1)  of  this  section. 

■  Lemma  Suppose  X:  D\  ->  R3  is  a  smooth  parametrized  surface  and  Y:  D2  —> 
R3  is  a  smooth  reparametrization  of  X  via  H:  D2  — >  D\,  where  we  denote  H(s,  f) 
by  (u,  v).  Then  the  standard  normal  vectors  Nx  and  NY  are  related  by  the  equation 

d(u,  v) 

NY(M)=!^-fNx(«,i;). 
a(s,  t) 


PROOF  First,  we  set  some  notation.  Since  Y  is  a  reparametrization  of  X  via  H, 
we  have,  from  Definition  2.3,  that 


Y(s,  0  =  X(H(s,  0)  =  X(ii,  u). 


(12) 


Write  (x(s,  t),  y(s,  t),  z(s,  f))  to  denote  Y(s,  t)  and  (x(u,  v),  y(u,  v),  z(u,  v))  to 
denote  X(m,  u),  even  though  this  is  a  small  abuse  of  notation. 
By  formula  (7)  of  §7. 1,  we  have 

xt  t    *     ^.z)  •     9(*>z)  .     d(x,  y) 
d(s,  t)        d(s,  t)        9(i,  f) 

If  we  apply  the  chain  rule  to  equation  (12),  we  obtain 

DY(s,  t)  =  DX(u,  v)DU(s,  t). 

Writing  out  this  matrix  equation,  we  get 


us  ut 

Vs  V, 


xs 

X, 

Xu  Xv 

ys 

yt 

y„  yv 

_  zs 

Zt  _ 

7.2  |  Surface  Integrals  487 


where  xs,  xt,  etc.  denote  partial  derivatives  of  the  component  functions  of  Y  and 
xu,  xv,  etc.  are  the  partial  derivatives  of  the  component  functions  of  X.  It  is  a 
matter  of  performing  the  matrix  multiplication  to  check  that 

[  xs   x,  ]  =  first  row  of  DY(s,  t) 

=  first  row  of  the  product  DX(u,  u)DH(j,  t) 
=  (first  row  of  DX(u,  v))  •  M(s,  t) 

—  [  Xy     Xy  ] 

Similar  results  hold  for  the  second  and  third  rows  of  DY(s ,  t).  We  may  recombine 
these  results  about  rows  and  establish  the  following  matrix  equations: 


X, 

Xu 

Xy 

Us 

ut 

y> . 

_  y„ 

yv  _ 

.  Vs 

vt  _ 

x, 

X, 

Xy 

Us 

ut 

zt  _ 

.  Zu 

Zv  _ 

v,  _ 

and 


Us  U[ 

vs  v, 


ys  yt 

yu 

yv 

Us  Ut 

_ZS    Zt  _ 

. Zu 

Zv  _ 

Taking  determinants,  we  find  that 


and 


d(x,  y) 

d(x,  y)  d(u,  v) 

3(5,0 

d(u,  v)  d(s,  t) 

d(x,  z) 

d(x,  z)  d(u,  v) 

3(s,t) 

d(u,  v)  d(s,  t) 

3(y.  z) 

d(y,z)  d(u,  v) 

d(s,t)  d(u,v)d(s,t) 
Thus,  returning  to  the  original  formula  for  NY,  we  find  that 


„  ,    .\      d(y>z)  d(u,  v). 

Nv  s,  t)  =  1  ■ 

'     3(ii,  v)  3(s,  t) 

d(u,  v) 


d(x,  z)  9(w,  v) .     d(x,  v)  d(u,  w), 

 J  H  K 

3(m,  v)  d(s,  t)       d(u,  v)  d(s,  t) 


d(s,  t) 


Nx(m,  v), 


as  desired. 


Proof  of  Theorem  2.4  We  use  Definition  2.1  and  the  change  of  variables 
theorem  for  double  integrals.  Thus,  by  Definition  2.1  and  the  lemma  just  proved, 


fifds 


J  £  f(Y(s,ty)\\NY(s,t)\\dsdt 


f Id2 


f(X(H(s,  0)) 


3(w,  u) 

3(5,  0 


|Nx(m(s,  0.  v(s,  0)11  ^Jfi??- 


488       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


From  the  change  of  variables  theorem,  it  follows  that 

I  I  fdS=  II  f(X(u,v))\\m.u,v)\\dudv=  II  fdS, 

J  Jy  J  J  Di  J  Jx 

by  Definition  2.1.  ■ 

Proof  of  Theorem  2.5  This  result  can  be  established  along  the  lines  of  the  pre- 
vious proof.  Beginning  with  Definition  2.2  and  using  the  lemma  just  established, 
we  have 


jj  F-dS  =  jj  F(Y(s,t))-~NY(s,t)dsdt 


(L 


F(X(HO,  0))  •  ^4  Nx("(^  0,  v(s,  t))ds  dt. 
D2  90, 0 


ds  dt, 


Therefore, 

JJ^F.dS  =  ±jjD  F(X(H(M)))-NX(«(M),KM)) 

where  we  take  the  "+"  sign  if  Y  is  an  orientation-preserving  reparametrization 
of  X  (since  the  Jacobian  d(u,  v)/d(s,  t)  is  positive  and  hence  equal  to  its  absolute 
value)  and  the  "— "  sign  if  Y  is  orientation-reversing.  By  the  change  of  variables 
theorem,  this  last  expression  is  equal  to 

±11  F(X(m,  w))-Nx(m,  v)du  dv  =  ±  1 1  F-dS, 

J  J  Di  J  JX 

by  Definition  2.2.  ■ 


7.2  Exercises 


1.  Let   X(s,t)  =  (s,s  +  t,t),    0<s<l,  0<t<2. 
Find 


(x2  +  y2+z2)dS. 


2.  Let  D  =  {(s,  t)  \  s2  +  t2  <  1,  s  >  0,  t  >  0}  and  let 
X:  D      R3  be  defined  by  X(s,  f)  =  (s  +  t,s-  t,  st). 

(a)  Determine  ffxf  dS,  where  f(x,  y,  z)  =  4. 

(b)  Find  the  value  of  ffx  F  •  dS,  where  F  =  xi  +  yj  + 
zk. 

3.  Find  the  flux  ofF  =  xi  +  3'j-|-zk  across  the  sur- 
face S  consisting  of  the  triangular  region  of  the  plane 
2x  —  2y  +  z  =  2  that  is  cut  out  by  the  coordinate 
planes.  Use  an  upward-pointing  normal  to  orient  S. 

4.  This  problem  concerns  the  two  surfaces  given  para- 
metrically  as 

X(s,  t)  =  (s  cos  t,  s  sinr,  3s2), 

0  <  s  <  2,  0  <  t  <  In. 


and 

Y(s,t)  =  (2s  cost,  2s  sin t ,  12s2), 
0  <  s  <  1, 0  <  t  <  An. 

(a)  Show  that  the  images  of  X  and  Y  are  the  same. 
(Hint:  Give  equations  in  x,  y,  and  z  for  the  sur- 
faces in  R3  parametrized  by  X  and  Y.) 

(b)  Calculate  / fx(y  i  -  x  j  +  z2  k)  •  dS  and  / fY(y  i  - 
x  j  +  z2  k)  •  dS.  Reconcile  your  answers. 

5.  Find  ffsx2dS,  where  S  is  the  surface  of  the  cube 
[-2,2]  x  [-2,2]  x  [-2,2]. 

6.  Find  / fs(x2  +  y2)  dS,  where  S  is  the  lateral  surface  of 
the  cylinder  of  radius  a  and  height  h  whose  axis  is  the 
z-axis. 

7.  Let  S  be  a  sphere  of  radius  a. 

(a)  Find  Jfs(x2  +  y2  +  z2)dS. 

(b)  Use  symmetry  and  part  (a)  to  easily  find  / fs  y2  dS. 


7.2  |  Exercises  489 


8.  Let  S  denote  the  sphere  x2  +  y2  +  z2  =  a2. 

(a)  Use  symmetry  considerations  to  evaluate  ffsxdS 
without  resorting  to  parametrizing  the  sphere. 

(b)  Let  F  =  i  +  j  +  k.  Use  symmetry  to  determine 
ffs  F  •  dS  without  parametrizing  the  sphere. 


9.  Let  S  denote  the  surface  of  the  cylinder  x2  +  y2 
—2  <  z  <  2,  and  consider  the  surface  integral 


4, 


(z-x2-y2)dS. 


(a)  Use  an  appropriate  parametrization  of  S  to  calcu- 
late the  value  of  the  integral. 

(b)  Now  use  geometry  and  symmetry  to  evaluate  the 
integral  without  resorting  to  a  parametrization  of 
the  surface. 

In  Exercises  10—18,  let  S  denote  the  closed  cylinder  with  bot- 
tom given  by  z  =  0,  top  given  by  z  =  4,  and  lateral  surface 
given  by  the  equation  x2  +  y2  =  9.  Orient  S  with  outward 
normals.  Determine  the  indicated  scalar  and  vector  surface 
integrals. 


I  j  zdS 


■  /// 


'•//, 


11.  /  /  ydS 


13.   /  /  xzdS 


14.  jj(xi  +  yj)-dS       15.  jj^zk-dS 

16.  j J  y3i-dS  17.  jj(-yi  +  xj)-dS 

» /pi.* 

In  Exercises  19-22,  find  the  flux  of  the  given  vector  field  F 
across  the  upper  hemisphere  x2  +  y2  +  z2  =  a2,  z  >  0.  Orient 
the  hemisphere  with  an  upward-pointing  normal. 

19.  F  =  v  j  20.  F  =  yi-x\ 

21.  F  =  —yi  +  x  j  -  k       22.  F  =  x2i  +  xy\  +  xzk 

23.  Let  S  be  the  parametrized  helicoid  X(s,  t  )  = 
(s  cos  t,  s  sinf,  f),  with  0  <  s  <  2,0  <  t  <  2n.  Deter- 
mine the  flux  ofF  =  yi  +  ;tj  +  z3k  across  S. 

24.  Let  F  =  2xi  +  2y\  +  z2k.  Find  ffsF-dS,  where  S 
is  the  portion  of  the  cone  x2  +  y2  =  z2  between  the 
planes  z  =  —2,  and  z  =  1,  oriented  with  outward- 
pointing  normal. 


25.  Find  the  flux  of  F  =  y3z  i  —  xy  j  +  (x  +  y  +  z)  k 
across  the  portion  of  the  surface  z  =  yex  lying  over 
the  unit  square  [0,  1]  x  [0,  1]  in  the. ty-plane,  oriented 
by  upward  normal. 

26.  Let  S  denote  the  tetrahedron  with  vertices  (0,  0,  0), 
(1,  0,  0),  (0,  2,  0),  (0,  0,  3)  oriented  by  outward  nor- 
mal, and  let  F  =  x2  i  +  4z  j  +  (y  —  x)  k.  Find  the  flux 
of  F  across  S. 

27.  Let  S  be  the  funnel-shaped  surface  defined  by  x2  + 


y 


z2  for  1  <  z  <  9  andx2  +  v2  =  1  forO  <  z  <  1. 


(a)  Sketch  S. 

(b)  Determine  outward-pointing  unit  normal  vectors 
to  S. 

(c)  Evaluate  ffs  F  •  dS,  where  F  =  —  yi  +  xj  +  zk 
and  S  is  oriented  by  outward  normals. 

28.  The  glass  dome  of  a  futuristic  greenhouse  is  shaped 
like  the  surface  z  =  8  —  2x2  —  2y2.  The  greenhouse 
has  a  flat  dirt  floor  at  z  =  0.  Suppose  that  the  tempera- 
ture T,  at  points  in  and  around  the  greenhouse,  varies 
as 

T(x,y,z)  =  x2  +  y2  +  3(z-2)2. 

Then  the  temperature  gives  rise  to  a  heat  flux  density 
field  H  given  by  H  =  —  k  V  T .  (Here  k  is  a  positive  con- 
stant that  depends  on  the  insulating  properties  of  the 
particular  medium.)  Find  the  total  heat  flux  outward 
across  the  dome  and  the  surface  of  the  ground  if  k  =  1 
on  the  glass  and  k  =  3  on  the  ground. 

29.  The  surface  given  by  X(s,  t)  =  (x(s,  t),  y{s,  t),  z(s,  t)), 
where 


(a  +  cos  -  sinf  —  sin  -  su^A  cos  s 
V  2  2  / 


2  2 

a  +  cos  -  sin  t  —  sin  -  sin  2f  ^  sin  s 

s  s 
sin  -  sin  t  +  cos  -  sin  2t 
2  2 


a  is  a  positive  constant,  and  0  <  s  <  2jt,  0  <  t  <  2jt, 
is  known  as  a  Klein  bottle. 

(a)  Use  a  computer  to  plot  this  surface  for  a  =  2. 

(b)  Determine  (and  describe)  the  s -coordinate  curve 
at  t  =  0. 

(c)  Calculate  the  standard  normal  vector  N  along  the 
i-coordinate  curve  at  t  =  0(i.e.,findN(.s,  0)).Note 
that  X(0,  0)  =  X(2jt,  0).  By  comparing  N(0,  0) 
and  N(2jt,  0),  comment  regarding  the  orientabil- 
ity  of  the  Klein  bottle.  (See  Example  8.) 


490 


Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


7.3   Stokes's  and  Gauss's  Theorems 

Here  we  contemplate  two  important  results:  Stokes's  theorem,  which  relates 
surface  integrals  to  line  integrals,  and  Gauss's  theorem,  which  relates  surface 
integrals  to  triple  integrals.  Along  with  Green's  theorem,  Stokes's  and  Gauss's 
theorems  form  the  core  of  integral  vector  analysis  and,  as  explained  in  the  next 
section,  can  be  used  to  establish  further  results  in  both  mathematics  and  physics. 

Stokes's  Theorem  

Stokes's  theorem  equates  the  surface  integral  of  the  curl  of  a  C1  vector  field  over  a 
piecewise  smooth,  orientable  surface  with  the  line  integral  of  the  vector  field  along 
the  boundary  curve(s)  of  the  surface.  Since  both  vector  line  and  surface  integrals 
are  examples  of  oriented  integrals  (i.e.,  they  depend  on  the  particular  orientations 
chosen),  we  must  comment  on  the  way  in  which  orientations  need  to  be  taken. 


DEFINITION  3.1  Let  S  be  a  bounded,  piecewise  smooth,  oriented  surface 
in  R3 .  Let  C  be  any  simple,  closed  curve  lying  in  S.  Consider  the  unit  normal 
vector  n  that  indicates  the  orientation  of  S  at  any  point  inside  C.  Use  n  to 
orient  C"  by  a  right-hand  rule,  so  that  if  the  thumb  of  your  right  hand  points 
along  n,  then  the  fingers  curl  in  the  direction  of  the  orientation  of  C.  (Equiva- 
lently,  if  you  look  down  the  tip  of  n,  the  direction  of  C  should  be  such  that  the 
portion  of  S  bounded  by  C  is  on  the  left.)  We  say  that  C  with  the  orientation 
just  described  is  oriented  consistently  with  S  or  that  the  orientation  is  the 
one  induced  from  that  of  S.  Now  suppose  the  boundary  dS  of  S  consists 
of  finitely  many  piecewise  C 1 ,  simple,  closed  curves.  Then  we  say  that  3  S 
is  oriented  consistently  (or  that  95  has  its  orientation  induced  from  that 
of  S)  if  each  of  its  simple,  closed  pieces  is  oriented  consistently  with  S. 


Some  examples  of  oriented  surfaces  with  consistently  oriented  boundaries 
are  shown  in  Figure  7.30.  If  the  orientation  of  5  is  reversed,  then  the  orientation  of 
3  S  must  also  be  reversed  if  it  is  to  remain  consistent  with  the  new  orientation  of  S. 

Now  we  state  a  rather  general  version  of  Stokes's  theorem,  a  proof  of  which 
is  outlined  in  the  addendum  to  this  section. 


Figure  7.30  Examples  of  oriented  surfaces 
and  curves  lying  in  them  having  consistent 
orientations.  On  the  right,  the  boundary  of  S2 
consists  of  three  simple,  closed  curves. 


7.3  I  Stokes's  and  Gauss's  Theorems  491 


THEOREM  3.2  (Stokes's  theorem)  Let  S  be  a  bounded,  piecewise  smooth, 
oriented  surface  in  R3.  Suppose  that  dS  consists  of  finitely  many  piecewise  C1, 
simple,  closed  curves  each  of  which  is  oriented  consistently  with  S.  Let  F  be  a 
vector  field  of  class  C1  whose  domain  includes  S.  Then 


V  x  F-dS  =  <f>  F  •  ds. 

J  as 


Figure  7.31  The 

paraboloid 
z  =  9  -  x2  -  y2 
oriented  with  upward 
normal  n.  Note  that 
the  boundary  circle 
C  is  oriented 
consistently  with  S. 


Theorem  3.2  says  that  the  total  (net)  "infinitesimal  rotation,"  or  swirling,  of 
a  vector  field  F  over  a  surface  S  is  equal  to  the  circulation  of  F  along  just  the 
boundary  of  S. 

EXAMPLE  1  Let  S  be  the  paraboloid  z  =  9  —  x2  —  y2  defined  over  the  disk 
in  the  xy-plane  of  radius  3  (i.e.,  S  is  defined  for  z  >  0  only).  Then  dS  consists  of 
the  circle 

C  =  {(x,y,z)\x2  +  y2  =  9,  z  =  0}. 

Orient  S  with  the  upward-pointing  unit  normal  vector  n.  (See  Figure  7.31.)  We 
verify  Stokes's  theorem  for  the  vector  field 


We  calculate 


V  x  F 


F  =  (2z-y)i  +  (x  +  z)j  +  (3x-2y)k. 


i  j  k 

d/dx       d/dy  d/dz 
2z  —  y      x  +  z      3x  —  2y 


=  (-2  -  l)i  +  (2  -  3)  j  +  (1  -  (-1)) k  =  -3i  -  j  +  2k. 

An  upward-pointing  normal  vector  N  is  given  by 

N  =  2xi  +  2yj  +  k. 

(This  vector  may,  of  course,  be  normalized  to  give  an  "orientation  normal"  n.) 
Therefore,  using  formula  (5)  of  §7.2  we  have,  where  D  =  {(x,  y)  \  x2  +  y2  <  9}, 

y^VxF-JS=  j  j  (-3i- j  +  2k)-(2xi  +  2yj  +  k)Jx^y 
(— 6x  —  2y  +  2)  dx  dy 


=  11  —6xdxdy—  11  2ydxdy  +  I  I  2dxdy. 

J  Jd  J  Jd  J  Jd 


By  the  symmetry  of  D  and  the  fact  that  —  6x  and  2y  are  odd  functions,  we  have 
that  the  first  two  double  integrals  are  zero.  The  last  double  integral  gives  twice 
the  area  of  D.  Thus, 


VxF-<iS  =  2-7r(32)=  18^. 


On  the  other  hand,  we  may  parametrize  the  boundary  of  S  as 
x  =  3  cos  t 

y  =  3  sin/      0  <  t  <  2n. 
z  =  0 


492       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Figure  7.32  The  surface 
S  =  {(*,  y,  z)  | 


Z  =  e 


-(*2+r) 


Z  >  l/e) 


has  boundary 

dS={(x,y,z)\ 
x2+y2  =  l,z=  l/e}. 


(This  parametrization  yields  the  orientation  desired  for  dS.)  Then 


(f  F-ds=  f 

Jss  Jo 


F(x(t))-x'(t)dt 


-2n 


=  f    (0  -  3  sin?,  3cosf  +  0,  9cosf  -  6sinf)- (-3  sin/,  3cosf,  0)dt 
Jo 


*2n 


f 

=  /     (9sin2f  +  9cos2f)^ 
Jo 


pin 


9dt  =  18tt, 


which  checks. 


EXAMPLE  2  Consider  the  surface  S  defined  by  the  equation  z  =  e  (x2+y2)  for 
z  >  l/e  (i.e.,  S  is  the  graph  of  f(x,  y)  =  e~(A'2+-y2)  denned  over  D  =  {(x,  y)  | 
x2  +  y2  <  1}).  Let 

F  =  (ey+z  -2y)i+  (xey+z  +  y)  j  +  ex+-Y  k. 

Then,  no  matter  which  way  we  orient  S,  we  can  see  that  JL  V  x  F  •  dS  looks  im- 
possible to  calculate.  Indeed,  suppose  we  take  the  upward-pointing  normal  vector 

N  =  2xe-(x2+y2)  i  +  2ye-(x2+y2)  j  +  k. 

Then,  because 

V  x  F  =  (ex+y  -  xey+z) i  +  (ey+z  -  ex+y)\  +  2k 
(you  may  wish  to  check  this),  using  formula  (5)  of  §7.2,  we  find  that 

//sV*F.„s 

=  jj    2xe-{x2+y2)(ex+y  -  xey+z)  +  2ye~^2+y2\ey+z  -  ex+y)  +  2]  dx  dy. 

We  will  not  attempt  to  proceed  any  further  with  this  calculation. 

It  is  tempting  to  use  Stokes's  theorem  at  this  point,  since  the  boundary  of  S  is 
the  circle  x2  +  y2  =  1,  z  =  l/e.  (See  Figure  7. 32.)  If  we  parametrize  this  circle  by 


X  =  cosf 
V  =  sinf 


1 


0  <  t  <  2jt, 


z  = 


then 


F  •  ds 


-L 
-I 


It, 


{e 


sin  t-\-\je 


2  sin  t ,  cos  t  esin '+ 1  /e  +  sin  t ,  ecos ,+sin ')  •  (-  sin  t ,  cos  t ,  0)  dt 


It, 


(2  sin2  t  -  sint  esint+1/e  +  cos2  tesint+1/e  +  cosf  sinf)^- 


Again,  we  have  difficulties. 

However,  the  power  of  Stokes's  theorem  is  that  if  S'  is  any  orientable 
piecewise  smooth  surface  whose  boundary  dS'  is  the  same  as  35  then,  subject  to 


7.3  I  Stokes's  and  Gauss's  Theorems  493 


5' 


n  dS =  dS' 

Figure  7.33  Both  S  and  5" 
have  the  same  boundary  and 
are  oriented  as  indicated. 
Therefore,  by  Stokes's  theorem, 
ffsVx*.dS  = 

fL  V  x  F  •  dS. 


orienting  S'  appropriately, 


F  •  ds  =  (b   F-ds  =        V  x  F  •  dS. 

IS  JdS  JdS1  J  JS' 

Hence,  we  may  evaluate  ffs  V  x  F  •  dS  by  using  a  different  surface!  (See  Fig- 
ure 7.33.) 

To  use  this  fact  to  our  advantage,  note  that  V  x  F  has  a  particularly  simple 
k-component.  Thus,  we  let  S'  be  the  unit  disk  at  z  —  l/e: 

S'  =  {(x,y,z)  |  x2  +  y2  <  l,z  =  l/e}. 

Consequently,  if  we  orient  5"  by  the  unit  normal  vector  n  =  +k,  we  have 


//IVxF-rfS  =  //, 
-//, 
-//. 


V  x  F  •  dS 


(V  x  F-n)dS 


2  dS  =  2  •  area  of  S'  =  2tx. 


Gauss's  Theorem  

Also  known  as  the  divergence  theorem,  Gauss's  theorem  relates  the  vector 
surface  integral  over  a  closed  surface  to  a  triple  integral  over  the  solid  region 
enclosed  by  the  surface.  Like  Stokes's  theorem,  Gauss's  theorem  can  assist  with 
computational  issues,  although  the  significance  of  the  result  extends  well  beyond 
matters  of  calculation. 

THEOREM  3.3  (Gauss's  theorem)  Let  D  be  a  bounded  solid  region  in  R3 
whose  boundary  3  D  consists  of  finitely  many  piecewise  smooth,  closed  orientable 
surfaces,  each  of  which  is  oriented  by  unit  normals  that  point  away  from  D.  (See 
Figure  7.34.)  Let  F  be  a  vector  field  of  class  C1  whose  domain  includes  D.  Then 


•—HI. 


V-FdV. 


Figure  7.34  A  solid  region  D  whose 
boundary  surfaces  are  oriented  so  that 
Gauss's  theorem  applies. 

By  a  closed  surface,  we  mean  one  without  any  boundary  curves,  like  a  sphere 
or  a  cube.  The  symbol  §  is  used  to  indicate  a  surface  integral  taken  over  a  closed 
surface  or  surfaces. 


494       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Si 


Figure  7.35  The 

solid  cylinder  D  of 
Example  3. 


Gauss's  theorem  says  that  the  "total  divergence"  of  a  vector  field  in  a  bounded 
region  in  space  is  equal  to  the  flux  of  the  vector  field  away  from  the  region  (i.e., 
the  flux  across  the  boundary  surface(s)). 

EXAMPLE  3  Let  F  be  the  radial  vector  field  x  i  +  y  j  +  z  k  and  let  D  be  the 
solid  cylinder  of  radius  a  and  height  b,  located  so  that  axis  of  the  cylinder  is  the 
z-axis  and  the  top  and  bottom  of  the  cylinder  are  at  z  =  b  and  z  —  0.  (See  Fig- 
ure 7.35.)  We  verify  Gauss's  theorem  for  this  vector  field  and  solid  region. 

The  boundary  of  D  consists  of  three  smooth  pieces:  (1)  the  bottom  surface  Si 
that  is  a  portion  of  the  plane  z  =  0  and  oriented  by  the  normal  iij  =  —  k,  (2)  the 
top  surface  S2  that  is  a  portion  of  the  plane  z  =  b  and  is  oriented  by  the  normal 
vector  112  =  k,  and  (3)  a  portion  of  the  lateral  cylinder  S3  given  by  the  equation 
x2  +  y2  =  a2  and  oriented  by  the  unit  vector  113  =  (x  i  +  y  j)/a.  (The  vector  113 
may  be  obtained  by  normalizing  the  gradient  of  f(x,  y,  z)  =  x2  +  y2  that  defines 
S3  as  a  level  set.)  Then 

F-dS  =  j  j  F-JS+  J  J  F-dS+  I  j  F-dS 

D  J  J  Si  J  JS2  J  Js3 

=  11  (xi  +  yj  +  zk)-(-k)dS+  /  /  (x  i  +  y  j  +  z  k)  •  kdS 

J  Js,  J  Js, 


+ 


J j^xi  +  yj  +  zk).  (^±H) 


x2  +  y2 


dS 


=  \l  -zdS+  j  j  zdS+  j  j 

J  JSi  J  Js2  J  JSi 

=  0+  [  f  bdS+  f  f  —dS, 

J  JS2  J  JSi  a 

since  along  Si ,  z  is  0;  along  52>  Z  is  equal  to  b;  and  along  S3 ,  x2  +  y2  =  a2.  Thus, 
(ft)    F  •  dS  =  b  •  area  of  S2  +  a  •  area  of  S3  =  bna2  +  ailnab)  =  3na2b, 

JJdD 

from  familiar  geometric  formulas. 
On  the  other  hand, 


V-F=  AW+  A(v)+l(z)  =  3, 
dx         ay  dz 


so  that 


which  can  be  checked  readily. 


3  d  V  =  3  •  volume  of  D  =  3na  b, 


In  general,  ifF  =  xi  +  yj  +  zk  and  D  is  a  region  to  which  Gauss's  theorem 
applies,  then 


§   F-dS=  HI  V-FdV=  ffl 

JJdD  J  J  Jd  J  J  J d 


3  dV  =  3  •  volume  of  D. 


Hence, 


(x  i  +  y  j  +  z  k)  •  dS  =  volume  of  D. 


3D 


7.3  I  Stokes's  and  Gauss's  Theorems  495 


Figure  7.36  The  union 
of  the  surfaces  S  and  S' 
enclose  a  solid  region  D 
to  which  we  may  apply 
Gauss's  theorem. 


Therefore,  we  may  use  surface  integrals  to  calculate  volumes  in  much  the  same 
way  that  we  used  Green's  theorem  to  calculate  areas  of  plane  regions  by  means 
of  suitable  line  integrals.  (See  §6.2,  especially  Examples  2  and  3.) 

EXAMPLE  4  Let 

F  =  ey  coszi  +  Vx3  +  1  sinzj  +  (x2  +  y2  +  3)k, 
and  let  S  be  the  graph  of 

z  =  (1  -  x2  -  y2)el-x2-iy2    for  z,  >  0, 

oriented  by  the  upward-pointing  unit  normal  vector.  It  is  not  difficult  to  see  that 
ffs  F  •  dS  is  impossible  to  evaluate  directly.  However,  we  will  see  how  Gauss's 
theorem  provides  us  with  elegant  indirect  means. 

Consider  the  piecewise  smooth,  closed  surface  created  by  taking  the  union  of 
S  and  S',  where  S'  is  the  portion  of  the  plane  z  =  0  enclosed  by  dS  (i.e.,  the  disk 
x2  +  y2  <  1,^  =  0).  Orient  S'  by  the  downward-pointing  unit  normal  z,  =  —  k  as 
shown  in  Figure  7.36.  Note  that  S  U  S'  forms  the  boundary  of  a  solid  region  D 
and  furthermore,  that  the  orientations  chosen  enable  us  to  apply  Gauss's  theorem. 
Doing  so,  we  have 

f  f  F-JS+  f  f  F-dS=  E   F-dS=  HI  V-¥dV. 

J  JS  J   ' S'  JJdD  J  J  J D 

Now,  it  is  a  simple  matter  to  check  that  V  •  F  =  0  for  all  (x,  y,  z).  Therefore,  the 
triple  integral  is  zero,  and  we  find  that 


so  that 


jj^¥-dS  +  jj^F-dS  =  0, 

JLF-ds=-fLF-ds- 


'S  J  JS' 

In  other  words,  because  F  is  divergenceless,  Gauss's  theorem  allows  us  to  replace 
the  original  surface  integral  by  one  that  is  considerably  easier  to  evaluate.  Indeed 
we  have 


F-dS 


f  j  F.(-k) dS 


j  j  (x2  +  y2  +  3)dx  dy, 


where  R  is  the  unit  disk  {(x,  y)  |  x2  +  y2  <  1}  in  the  plane.  Now,  we  switch  to 
polar  coordinates  to  find 

1 1  (x2 +  y2 +  3)dxdy  =  f     f  (r2  +  3)rdrd0 
J  Jr  Jo  Jo 


=  (y+\r2)\i_ 

Jo 
Jo 


de 


EXAMPLE  5   Consider  the  vector  field 

=    xi  +  yj  +  zk 

~  (x2  +  yi  +  z2W 


496       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


(This  is  an  example  of  an  inverse  square  vector  field.)  The  flux  across  the  sphere 
x2  +  y2  +  z2  =  a2  oriented  by  outward  unit  normal  n  is  given  by  §s  F  •  dS  = 
§s  F  •  n  dS.  The  unit  normal  to  S  may  be  computed  as 


n  = 


V(x2  +  y2  +  z2)        2xi  +  2yj+2zk      xi  +  yj  +  zk 


||V(x2  +  y2  +  z2)||      ^4x2  +  4y2  +  4z2 


since  x2  +  y2  +  z2  =  a2  on  the  surface  of  the  sphere.  In  a  similar  way,  we  may 
write  F(x,  y,  z)  as  (x  i  +  y  j  +  z  k)/a3  whenever  (x,  y,  z)  is  a  point  on  the  sphere. 
Hence, 

'xi  +  yj  +  zk\    /xi+yj  +  zkN 


F-dS 


F-ndS 


x2  +  y2  +  z2 


dS 


1 


-(surface  area  of  S) 

,  cr  }  a1 

=  ^{4na2)  =  4jt. 

Now  we  note  that  the  partial  derivative  of  the  first  component  of  F  with 
respect  to  x  is 

dFi  _  (x2  +  y2  +  z2)3/2  -  3x2(x2  +  y2  +  z2)1/2       -2x2  +  y2  +  z2 
3x 
Similarly, 


(x2  +  y2  +  z2)3 


(x2  +  y2  +  z2)5/2 ' 


9f2 

3y 


2y2  +  z2 


(x2 


z2f/2 


and 


3F3 


+  y2  -  2z2 


(x2  +  y2  + 


z2)5/2- 


Thus,  V  •  F  =  0,  and  so  any  triple  integral  of  V  •  F  must  be  zero.  This  would  seem 
to  be  at  odds  with  Gauss's  theorem.  There  is  no  contradiction,  however.  Note  that 
the  vector  field  F  is  not  defined  at  the  origin.  Therefore,  the  hypothesis  that  F 
be  defined  throughout  the  solid  region  enclosed  by  S  is  not  satisfied  and,  hence, 
Gauss's  theorem  does  not  apply  in  this  situation.  ♦ 


Figure  7.37  A 

sphere  of  radius  a 
used  to  understand 
the  divergence  of  a 
vector  field. 


The  Meaning  of  Divergence  and  Curl   

Part  of  the  significance  of  Stokes's  theorem  and  Gauss's  theorem  is  that  they 
provide  a  way  to  understand  the  meaning  of  the  divergence  and  curl  of  a  vector 
field  apart  from  the  coordinate-based  definitions  of  §3.4.  By  way  of  explanation, 
we  offer  the  following  two  results: 

PROPOSITION  3.4  Let  F  be  a  vector  field  of  class  C1  in  some  neighborhood  of 
the  point  P  in  R3.  Let  Sa  denote  the  sphere  of  radius  a  centered  at  P,  oriented 
with  outward  normal.  (See  Figure  7.37.)  Then 

divF(P)=  lim  (j^  F-dS. 


>0+  47Tfl3 


PROOF  We  have,  by  Gauss's  theorem,  that 


lim 


>o+  4na! 


F-dS=  lim 


3 


fl-»-0+  47T(33 


divFdV, 


(1) 


7.3  I  Stokes's  and  Gauss's  Theorems  497 


C„  =  dS„ 


where  Da  is  the  solid  ball  of  radius  a  enclosed  by  Sa.  Next,  we  use  a  result 
known  as  the  mean  value  theorem  for  triple  integrals,  which  states  that  if  / 
is  a  continuous  function  of  three  variables  and  D  is  a  bounded  connected  solid 
region  in  space,  then  there  is  some  point  Q  6  D  such  that 


///. 


f(x,y,z)dV  =  /(g).  volume  of  D. 

D 

In  our  present  situation,  this  result  implies  that  there  must  be  some  point  Q  e  Da 
such  that 


///. 


divFJV  =  divF(fi).  (volume  of  Da)  =  (f na3)  divF(g).  (2) 
Applying  formula  (2)  to  formula  (1),  we  have 

lim  $  F-dS  =  lim  divF(g)  =  divF(P), 


a->o+  Ana}  Jfs  a-*o+ 


since,  as  a  ->  0+,  the  ball  Da  becomes  smaller,  "crushing"  Q  onto  P.  ■ 

PROPOSITION  3.5  Let  F  be  a  vector  field  of  class  C1  in  a  neighborhood  of  the 
point  P  in  R3.  Let  Ca  be  the  circle  of  radius  a  centered  at  P  situated  in  the  plane 
containing  P  that  is  perpendicular  to  the  unit  vector  n.  (See  Figure  7.38.)  Then 
the  component  of  curl  F(P)  in  the  n-direction  is 


n-curlF(P)=  lim  — -  <f>  F-ds, 

a->-o+  no1  JCa 

Figure  7.38  A  circle  of  where  Ca  is  oriented  by  a  right-hand  rule  with  respect  to  n. 
radius  a  centered  at  P.   


PROOF  Let  Sa  denote  the  disk  of  radius  a  in  the  plane  of  Ca  enclosed  by  Ca .  By 
Stokes's  theorem, 

lim  — l—  S)  F  •  ds  =  lim  — —  f  f  VxF-JS 

£i^o+  7ta2  Jc  a^o+  na2  J  Js 


.0+ 


J  Jsu 


=  lim  — -  /  /  (VxF.n)iJS.  (3) 


There  is  a  mean  value  theorem  for  surface  integrals  (similar  to  the  mean  value 
theorem  for  triple  integrals  used  in  the  proof  of  Proposition  3.4)  enabling  us  to 
conclude  that  there  must  be  some  point  Q  in  Sa  for  which 


( V  x  F  •  n)  dS  =  (V  x  F(g)  •  n)(area  of  Sa) 

=  na2(V  xF(G)-n),  (4) 
since  Sa  is  a  disk  of  radius  a.  Therefore,  using  equations  (3)  and  (4),  we  find  that 


1     f  _   .  1 


lim  — -  A  F-ds=  lim  — -(jra2V  x  F(Q)  •  n) 

a->-0+  na2  JCa  a^0+  na2 

=  lim  V  x  F(Q)-n 
=  V  x  F(P)  •  n, 

since  Q  — >•  P  as  a  —>  0+ . 


498       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Propositions  3.4  and  3.5  justify  our  claims,  made  in  §3.4,  about  the  intuitive 
meanings  of  the  divergence  and  curl  of  a  vector  field.  The  quantity  §s  F  •  dS 
used  in  Proposition  3.4  is  the  flux  of  F  across  the  sphere  Sa,  and  so 


lim 


>0+  47Tfl3 


F  •  dS 


n  =  k 


Figure  7.39  The 

configuration  needed 
to  calculate  the 
k-component  of 
curl  F(xo),  using 
Proposition  3.5. 


is  precisely  the  limit  of  the  flux  per  unit  volume,  or  the  flux  density  of  F  at  P. 
Similarly, 

1 


lim 

a^0+  Ttaz 


F  •  ds 


is  the  limit  of  the  circulation  of  F  along  Ca  per  unit  area,  or  the  circulation 
density  of  F  at  P  around  n.  In  particular,  Proposition  3.5  shows  that  curlF(P) 
is  the  vector  whose  direction  maximizes  the  circulation  density  of  F  at  P  and 
whose  magnitude  is  equal  to  the  circulation  density  around  that  direction  (or  else 
curlF(ip)  is  0  if  the  circulation  density  is  zero). 

In  fact,  we  can  turn  our  approach  to  divergence  and  curl  completely  around 
and,  instead  of  defining  the  divergence  and  curl  by  means  of  coordinates  and  the 
del  operator  and  proving  Propositions  3.4  and  3.5,  use  the  surface  and  line  integral 
formulas  of  Propositions  3.4  and  3.5  to  define  divergence  and  curl  and  derive  the 
coordinate  formulations  from  the  limiting  integral  formulas. 

Write  F  as 

M(x,  y,z)i  +  N(x,  y,  z)  j  +  P(x,  y,  z)  k, 

where  M,  N,  and  P  are  functions  of  class  C1  in  a  neighborhood  of  the  point 
xo  =  (xq,  yo,  Zo)-  We  will  demonstrate  how  to  recover  the  coordinate  formula 


(dP 

V  x  F  =  — 
\dy 


dN 

~dz 


i  + 


dM 

~dz 


dP 

dx 


j  + 


dN  dM\ 
 k 

dx       dy  J 


from  the  formula  in  Proposition  3.5.  (A  similar  argument  can  be  made  to  derive 
the  coordinate  formula  for  the  divergence,  the  details  of  which  we  leave  to  you.) 
The  idea  is  to  let  the  unit  vector  n  equal,  in  turn,  i,  j,  and  k  in  the  formula  in 
Proposition  3.5  and  thereby  to  determine  the  components  of  the  curl. 

First,  let  n  =  k,  so  that  Ca  is  the  circle  of  radius  a  in  the  horizontal  plane 
z  =  zo,  oriented  counterclockwise  around  k.  (See  Figure  7.39.)  Then  Ca  may  be 
parametrized  by 


x  =  xq  +  a  cos  / 
y  =  yo  +  a  sin  t 

Z  =  ZO 


0<t<2n. 


Therefore, 


&  F-ds=  f  F(x(t))-x'(t)dt 

jc„  Jo 


/•2tt 

=  /  0 

JO 


(M(x(f))i  +  N(x(t))j  +  P(x(t))k)-(-a  sinr  i  +  a  cos  t \)dt 


•2ti 


=  a  f  (-sin?  M(x(t))  +  cost  N(x(t)))  dt. 
Jo 


7.3  I  Stokes's  and  Gauss's  Theorems  499 


Next,  use  Taylor's  first-order  formula  (see  §4.1)  on  M  and  N ,  which  yields,  near 
x0  (i.e.,  for  (x,  y,  z)  «  (x0,  yo,  Zo)), 

M(x,  y,  z)  «  M(xo)  +  Mx(xo)(x  -  x0)  +  Mv(x0)(y  -  y0)  +  Mz(x0)(z  -  zo); 

JV(x,  y,  z)  «         +  W*(x0)(x  -  x0)  +  Ny(x0)(y  -  yo)  +  N:(x0)(z  -  Zo). 

Along  the  small  circle  Ca,  we  have  x  —  xq  =  a  cos/,  y  —  yo  =  a  sin?,  and  z  — 
zo  =  0,  so  that,  using  the  approximations  for  M  and  TV,  we  have 

/>2ir 

F-ds^a  /     —  sin ?  [M(x0)  +  Mx(x0)a  cos?  +  My(x0)a  sin? ]  dt 
Jo 


+ 


a  I  i 

Jo 


cos  t [N(xo)  +  Nx(xo)a  cos  ?  +  Ny(xo)a  sin  ?  ]  dt 


'  / 

Jo 


=  — aM(xo)  /     sin?  dt  —  a  MT(xo)  /     sin?  cos?  J? 


2tt 


p27T 

'  / 

Jo 


a  M,(xo)  /     sin  ?J?  +  aiV(xo)  /     cos?  J? 


Jo 


-In 


+ 


nLlt  nllt 

a2Nx(x0)  I     cos2  t  dt  +  a2  Ny(x0)  I     sin?  cos?  J?.  (5) 
Jo  Jo 

This  last  equality  holds  because  M(x0),  MA  (x0),  etc.,  do  not  involve  ?  and  so  may 
be  pulled  out  of  the  appropriate  integrals.  You  can  check  that 


piit  pin  t> 

I  smt  dt  =  J  cos?  dt  =  I 
Jo  Jo  Jo 

/     sin2  tdt  = 
Jo  Jo 


sin?  cost  dt  =  0, 


COS   t  dt  =  71. 


Therefore,  the  approximation  in  (5)  simplifies  to 


F  •  ds  «  -na2My(x0)  +  jra2Nx(x0). 


Now,  the  error  involved  in  the  approximation  for  j>c  F-ds  tends  to  zero  as  Ca 
becomes  smaller  and  smaller.  Thus, 

k •  curl F(x0)  =  lim  — *-r  £>  F-ds 

a->-o+  nal  JCa 

=  lim  (-Mv(x0)  +  /V,(x0)) 

=  Nx(x0)  -  My(x0). 

If  we  let  n  =  j,  so  that  Ca  is  the  circle  parametrized  by 

x  =  xi)  +  a  sin  ? 
y  =  y0  0  <  ?  <  lit, 

z  =  Zo  +  a  cos  t 

then  a  very  similar  argument  to  the  one  just  given  shows  that 

j-curlF(x0)  =  M,(xo)-JPl(x0). 


Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


Finally,  let  n  =  i,  so  that  Ca  is  parametrized  by 


x  =  x0 

y  =  yo  +  a  cos  t     0  <  t  <  2n, 
z  =  zo  +  a  sin  t 

to  find 

i  •  curlF(xo)  =  Py(xo)  -  Nz(x0). 
We  see  that  the  i-,  j-,  and  k-components  of  the  curl  of  F  are  as  stated  in  §3.4. 

Addendum:  Proofs  of  Stokes's  and  Gauss's  Theorems 
Proof  of  Stokes's  theorem 

Step  1.  We  begin  by  establishing  a  very  special  case  of  the  theorem,  namely, 
the  case  where  the  vector  field  F  =  M(x,  y,  z)  i  (i.e.,  F  has  an  i-component  only) 
and  where  the  surface  5  is  the  graph  of  z  =  f(x,  y),  where  /  is  of  class  C1  on  a 
domain  D  in  the  plane  that  is  a  type  1  elementary  region.  (See  Figure  7.40.)  To 
be  explicit,  the  region  D  is 

{(x,  y)  |  y(x)  <  y  <  S(x),  a  <x  <b), 

where  y  and  S  are  continuous  functions.  We  assume  that  S  is  oriented  by  the 
upward-pointing  unit  normal. 

y 


(D 

^=5(x). 

D 

® 

® 

y  =  j(x) 

a 

b 

Top  view 

Figure  7.40  The  surface  S  is  the  graph  of  f(x,  y)  for  (x,  y)  in 
the  type  1  region  D  shown  at  the  right. 

First,  evaluate  §as  F  •  ds.  The  boundary  3S  consists  of  (at  most)  four  smooth 
pieces  parametrized  as  follows.  (The  subscripted  curves  correspond  to  the  encir- 
cled numbers  in  Figure  7.40.) 


C 


l  • 


x  =  t 

v  =  y(t)  a  <t  <  b; 

z  =  fit,  y(t)) 


Co 


x  =  b 

V  =  t  y(b)  <  t  <  8(b); 

Z  =  f(b,  t) 


C3  : 


x  =  t 

y  =  8(t)  a  <t  <b; 

z  =  fit,  S(t)) 


x  =  a 

y  =  t  y(a)  <  t  <  8(a). 

z  =  f(a,  t) 


The  parametrizations  shown  for  C3  and  C4  induce  the  opposite  orientations  to 
those  indicated  in  Figure  7.40.  Therefore, 


(f  F-ds=  j   F.Js+/   F-d$-  I  F-ds-  j 

JdS  JCi  JC2  Jc3  JC4 


F-ds. 


7.3  I  Stokes's  and  Gauss's  Theorems  501 


Consider  the  integral  over  C\ .  Since  F  has  only  an  i-component, 

f  F-ds=  f  Mdx  +  0dy  +  0dz=  f  M(t,  y(t),  f(t,  y(t)))dt. 

Jd  JC,  Ja 

The  line  integral  fc  F  •  ds  may  be  calculated  in  a  manner  similar  to  that  for 
fc  F  •  ds.  In  particular,  we  obtain 

f  ¥-ds=  f  M(t,S(t),  f(t,8(t)))dt. 

JC}  J  a 

For  the  integral  over  C2,  note  that  x  is  held  constant.  Thus, 


/   F  •  ds  =  f  M  dx  =  0. 

Jc2  Jc2 

Likewise,  fc  F  •  ds  =  0.  The  result  is 
<f  F.ds=  f  M(t,y(t),f(t,y(t)))dt-f  M(t,  8(t),  f(t,  S(t)))dt 

J  dS  J  a  J  a 

=  f  [M(x,y(x),f(x,y(x)))-M(x,S(x),f(x,S(x)))]dx.  (6) 

J  a 

(In  this  last  equality  we've  made  a  change  in  the  variable  of  integration.) 

Now  we  compare  the  line  integral  to  the  surface  integral  ffs  V  x  F  •  dS.  For 
F  =  M(x,  y,z)i,  we  have  V  x  F  =  Mzj  —  My  k,  so  formula  (5)  of  §7.2  yields 

J jvxV.dS  =  j jjMzi-MyK).(-fxi- fy\+k)dxdy 

J  a    J y( 


*b  r*W  /    dM  df  dM\ 


ly(x) 

The  chain  rule  implies 


)  dydx. 
dz  dy       dy  ) 


3  dM     dM  df 

—  (M(x,  y,  f(x,  y))  =  —  +  —  -f. 
dy  dy       dz  dy 

Thus,  using  the  fundamental  theorem  of  calculus,  we  have 

,  ,  *b  ,«(*)  a 

VxF-dS=       /  (M(x,y,  f(x,y)))dydx 

J  Js  Ja    Jy(x)  3y 


=  f  -M(x,y,f(x,y))\yyZSy%dx 

J  a 


=  f  [-M(x,S(x),  f(x,S(x)))  +  M(x,y(x),  f(x,y(x)))]dx, 

J  a 

which  agrees  with  equation  (6). 

We  may  readily  extend  this  result  to  surfaces  that  are  graphs  of  z  =  f(x,  y), 
where  the  point  (x,  y)  varies  through  an  arbitrary  region  D  in  the  xy-plane  via  a 
two-stage  process:  First,  establish  the  result  for  regions  D  that  may  be  subdivided 
into  finitely  many  elementary  regions  of  type  1  and  then  apply  a  limiting  argument 
similar  to  the  one  outlined  in  the  proof  of  Green's  theorem  in  §6.2. 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Figure  7.41  The  surface  S 
is  a  portion  of  the  plane 
x  =  c.  If  F  =  Mix,  y,  z)i, 
then  F  is  always 
perpendicular  to  S,  in 
particular,  to  any  vector  T 
tangent  to  9S. 


Figure  7.42  The  surfaces  Si 
and  S2  share  the  curve  C  as 
part  of  their  boundaries.  C 
receives  one  orientation  as 
part  of  9 Si,  and  the  opposite 
orientation  as  part  of  dSz- 


Step  2.  Still  keeping  F  =  Mi,  note  that  the  argument  given  in  Step  1  works 
equally  well  for  surfaces  of  the  form  y  =  fix,  z) — simply  exchange  the  roles  of 
y  and  z  throughout  Step  1 . 

It  is  also  not  difficult  to  see  that  if  S  is  a  portion  of  the  plane  x  =  c  (where 
c  is  a  constant),  then  Stokes's  theorem  for  F  =  Mi  holds  for  this  case,  too.  On 
the  one  hand  n  =  ±i  for  such  a  plane  (depending  on  the  orientation  chosen)  and 
V  x  F  =  Mz  j  —  My  k,  as  we  have  seen.  Hence, 


/  /  VxF-JS=  /  /  (VxF-n)JS  =  /  /  0 
J  Js  J  Js  J  Js 


dS  =  0. 


On  the  other  hand  since  F  has  only  an  i-component,  F  is  always  parallel  to 
n  and  thus  perpendicular  to  any  tangent  vector  to  S,  including  vectors  tangent 
to  any  boundary  curves  of  S.  Therefore,  <fas  F  •  ds  =  0  also.  (See  Figure  7.41.) 
Now,  suppose  that  S  =  Si  U  S2,  where  Si  and  S2  are  each  one  of  the  graph  of 
z  =  fix,  y),  the  graph  of  y  =  fix,  z),  or  a  portion  of  the  plane  x  =  c.  Assume 
that  Si  and  S2  coincide  along  part  of  their  boundaries,  as  shown  in  Figure  7.42. 
The  surfaces  Si  and  S2  inherit  compatible  orientations  from  S.  If  C  denotes 
the  common  part  of  3  Si  and  3S2,  then  we  may  write  3  Si  as  C\  U  C  and  3S2  as 
C2  U  C,  where  C\  and  C2  are  disjoint  (except  at  their  endpoints).  If  3S,  is  oriented 
consistently  with  S,  for  i  =  1,2,  then  note  that  C  will  be  oriented  one  way  as  part 
of  3  Si  and  the  opposite  way  as  part  of  3S2.  From  this  point,  let  us  agree  that  C 
denotes  the  curve  oriented  so  as  to  agree  with  the  orientation  of  3 Si. 

Now  Stokes's  theorem  with  F  =  Mi  holds  on  both  Si  and  S2;  on  Si  we  have 


/  /  VxF-dS=  (p   F •  ds  =  I  ¥-ds+  j  ¥■ 

whereas  on  S2  we  have 

/  /  V  x¥-dS=  (t   F •  ds  =  I  F-ds  -  I  F 

J  Js?  J  as,  Jc?  Jc 


ds, 


(7) 


(8) 


in  view  of  the  remarks  made  regarding  the  orientation  of  C.  Now,  we  consider 
S  =  Si  U  S2,  noting  that  C  is  not  part  of  3S.  We  see  that 


ffvxF-dS=ff  VxF-dS+ff 

J  JS  J  JSi  J  J s2 

Using  equations  (7)  and  (8)  and  canceling,  we  find  that 


V  x  F-dS. 


/   F-ds+  F-ds 

JCi  Jc2 


F-ds. 


us 


Thus,  Stokes's  theorem  holds  in  this  case,  or,  indeed,  in  the  case  where  S  can  be 
written  as  a  finite  union  Si  U  S2  U  •  •  •  U  S„  of  the  special  surfaces  just  described. 

From  a  practical  point  of  view,  any  particular  surfaces  you  are  likely  to 
encounter  will  be  decomposable  as  finite  unions  of  the  special  types  of  surfaces 
previously  described.  However,  not  all  piecewise  smooth  surfaces  are,  in  fact,  of 
this  form.  So  to  finish  a  truly  general  proof  of  Stokes's  theorem  when  F  =  Mi, 
some  further  limit  arguments  are  needed,  which  we  omit. 

Step  3.  Finally,  by  permuting  variables,  we  essentially  repeat  Steps  1  and  2  in 
the  cases  where  F  =  N(x,  y,  z)  j  or  F  =  P(x,  y,  z)k.  In  general,  by  the  additivity 


7.3  I  Stokes's  and  Gauss's  Theorems  503 


of  the  curl,  we  have  that 
V  x  F  •  dS 


j xF-dS  =  jjvx(Mi  +  Nj  +  Pk)-dS 


=  j jv  x(Mi)-dS  +  JJvx(Nj)-dS 

+  j  x(Pk)-dS. 
Using  the  versions  of  Stokes's  theorem  just  established,  we  see  that 

//VxF-dS=(f>  Mi-ds+f  N]-ds+(b  Pk-ds, 

J  Js  JdS  JdS  Jas 

=  <b  (Mi  +  N\  +  Pk)-ds=  f  F-ds, 

Jas  Jas 

as  desired. 


Figure  7.43  The  type  1  region 
D  and  its  shadow  region  R  in  the 
xy-plane. 


Proof  of  Gauss's  theorem 

Step  1.  We  prove  a  very  special  case  of  Gauss's  theorem,  namely,  the  case  in 
which  F  =  P  k  (where  P(x,  y,  z)  is  of  class  C1  on  a  domain  that  includes  the 
solid  region  D)  and  where  D  is  an  elementary  region  of  type  1,  as  in  Figure  7.43. 
We  denote  the  bottom  surface  boundary  of  D  by  S\  and  take  it  to  be  given  by 
the  equation  z  =  (p(x,  y),  where  cp  is  of  class  C1  and,  similarly,  we  let  52  denote 
the  top  surface  boundary  and  assume  it  is  given  by  the  equation  z  =  is(x,  y), 
where  ^  is  also  of  class  C1 .  The  lateral  surface  boundary  is  denoted  by  S3 ;  it  may 
reduce  to  a  curve  or  be  empty  but  otherwise  is  a  cylinder  over  the  boundary  curve 
of  a  region  R  in  the  plane  forming  the  shadow  of  D. 

Orienting  3  D  =  Si  U  52  U  S3  with  outward-pointing  normal  vectors,  we  have 


F  •  dS 


f  f  F-dS+  f  f  F-dS+  f  f 

J  Jsi  J  Js7  J  Js-t 


F-dS. 


The  orientation  normal  to  S\  should  be  downward-pointing,  hence  parallel  to 
<pxi  +  <py\  —  k,  the  opposite  of  the  normal  vector  obtained  from  the  standard 
parametrization  of  Si.  Therefore,  using  formula  (5)  of  §7.2,  we  have 

/  /  F-dS=  /  /  P(x,y,(p(x,y))k-((pxi  +  <py]-k)dxdy 

J  JS,  J  J R 


-JL 


P(x,  y,  <p{x,  y))dx  dy. 


Similarly,  the  orientation  normal  to  S2  should  be  upward-pointing,  and  so 
//  F-dS=  if  P(x,y,f(x,y))-(-irxi-fy}+K)dxdy 

J  J  Si  J  J  R 

=  j  j  P(x,  y,  \j/(x,  y))dx  dy. 

Now  the  lateral  surface  S3,  if  it  is  nonempty,  is  a  cylinder  over  a  curve  in  the 
xy-plane.  Hence,  S3  is  defined  by  one  or  more  equations  of  the  form  g(x ,  y)  =  c. 


504       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Figure  7.44  The  regions  D\  and 
D2  share  the  surface  S  as  part  of 
their  boundaries.  This  common 
surface  S  inherits  one  orientation 
as  part  of  dD\  and  the  opposite 
orientation  as  part  of  d Di- 


li follows  that  any  normal  vector  to  53  can  have  no  k-component.  Thus, 

/  /  F-dS  =  I  I  (Pk-n3)J5=  /  /  0dS  =  0. 

J  JSx  J  Js%  J  Js% 


IS3  J  JS3 

Putting  these  results  together,  we  find  that 


F-dS  =  f  f 

D  J  JR 


[P(x,  y,  f{x,  y))  -  P(x,  y,  <p(x,  y))]dxdy. 


On  the  other  hand,  if  F  =  Pk,  then  V  •  F  =  dP/dz,  so  that,  by  the  funda- 
mental theorem  of  calculus, 


r  r  r  r  r  r^'y)  dp 

/  /  /  V-FJV=  /  /  /  — dzdxdy 
J  J  Jd  J  Jr  Jip(x.v)  9z 


-//. 


tp(x,y) 

[P(x,  y,  x[r(x,  y))  -  P(x,  y,  cp(x,  y))]dxdy. 


Therefore,  Gauss's  theorem  holds  in  this  special  case. 

Step  2.  We  may  repeat  Step  1  for  F  =  Mi  and  D  an  elementary  region  of 
type  2,  and  for  F  =  A^j  and  D  of  type  3.  If  D  is  an  elementary  region  of  type  4 
(meaning  D  is  simultaneously  of  types  1,2,  and  3),  then 


F  •  dS 


3D 


(Mi  +  N]  +  PkWS 


i)D 


Mi-dS  + 


•  ID 


Nj-dS+Qp  Pk-dS 

3D  JJdD 


=HLv-M,dv+UL 
-ffL 

-ffL 


!  V-NjdV  + 
V  .(Mi+  Nj  +  Pk)dV 
V-FdV. 


ffL 


V-PkdV 


Step  3.  Suppose  that  D  =  D\  U  D2,  where  each  of  D\  and  D2  is  a  type  4 
elementary  region,  and  that  D\  and  Do  coincide  just  along  part  of  their  boundaries 
as  shown  in  Figure  7.44.  Let  S  denote  the  common  part  of  dD\  and  dD2-  Then 
we  write  dD\  as  Si  U  S  and  dD2  as  S2  U  S,  where  S\  and  S2  are  disjoint  (except 
perhaps  along  portions  of  their  respective  boundaries).  If  we  orient  dD\  and  dD2 
with  outward  normals  so  as  to  satisfy  the  hypotheses  of  Gauss's  theorem,  then  S 
will  be  oriented  one  way  as  part  of  dDi  and  the  opposite  way  as  part  of  dD2.  Let 
us  agree  that  the  symbol  S  denotes  the  surface  oriented  so  as  to  agree  with  the 
orientation  of  3  D\ . 

Applying  Gauss's  theorem  to  both  D\  and  D2,  we  obtain 


ffL 
ffL 


V-FdV 


S7-FdV 


F-dS 


SD, 


F-dS 


8D2 


i  i Si  i  is 


F  •  dS 


F-dS. 


7.3  |  Exercises  505 


7.3  Exercises 


Combining  these  last  two  equations,  we  find  that 

/  /  /  V-FdV  =  iff  V  -FdV  +  /  /  /  V-FdV 

J  J  JD  J  J  JDi  J  J  J D2 

=  f  f  F-dS  +  f  f  F-dS 

J  J  Si  J  Js2 

F-dS, 


3D 


since  3D  =  S\  U  S2. 


Step  4.  The  result  of  Step  3  may  be  extended  to  regions  D  that  can  be 
decomposed  as  a  union  of  finitely  many  type  4  regions.  However,  not  all  regions 
to  which  Gauss's  theorem  applies  meet  this  criterion.  Consequently,  to  finish  a 
truly  general  proof,  we  once  again  need  an  argument  using  suitable  limits  of 
regions  and  their  integrals,  which  we  omit.  ■ 


In  Exercises  1—4,  verify  Stokes  s  theorem  for  the  given  surface 
and  vector  field. 

1.  S  is  defined  by  x2  +  y2  +  5z  =  1,  z  >  0,  oriented  by 
upward  normal; 

F  =  xzi  +  yz)  +  (x2  +  y2)k. 

2.  S  is  parametrized  by  X(s,  t)  =  (s  cost,  s  sinf,  f),  0  < 
s  <  1,  0  <  t  <  jt/2; 

F  =  zi  +  xj  +  yk. 

3.  S  is  defined  by  x  =  yl6  —  y1  —  z1', 

F  =  xi  +  yj  +  zk. 

4.  S  is  defined  by  x2  +  y2  +  z2  =  4,  z  <  0,  oriented  by 
downward  normal; 

F  =  (2y  -  z)  i  +  (x  +  y2  -  z)  j  +  (4y  -  3x)k. 

5.  Let  S  be  the  "silo  surface,"  that  is,  5  is  the  union  of 
two  smooth  surfaces  S\  and  S2,  where  Si  is  defined  by 

x2  -\-  y2  =  9,     0  <  z  <  8 

and  S2  is  defined  by 

x2  +  y2  +  (z  -  8)2  =  9,    z  >  8. 

Find  ffs  V  x  F  •  dS,  where 

F  =  (jc3  +  xz  +  yz2)  i  +  (xyz3  +  y1)  j  +  x2z5  k. 

In  Exercises  6—9,  verify  Gauss  s  theorem  for  the  given  three- 
dimensional  region  D  and  vector  field  F. 

6.  F  =  *i+yj  +  zk, 

D  =  {(x,  y,  z)  I  0  <  z  <  9  -  x2  -  y2} 

7.  F  =  (y  —  x)  i  +  (y  —  z)  j  +  (x  —  y)  k,  D  is  the  unit 
cube  [0,  1]  x  [0,  1]  x  [0,  1] 


8.  F  =  x2i+  vj  +  zk,  D  =  {(x,y,z)  |  x2  +  y2  +  1  < 
z  <  5} 

9.  F=  -y,i7+yj?+Zk,  D  =  {(x,y,z)\a2<x2  + 

jx2  +  y2  +  z2 

y2  +  z2<  b2} 

10.  Verify  that  Stokes's  theorem  implies  Green's  theo- 
rem. (Hint:  In  Stokes's  theorem  take  F(x,  y,  z)  = 
M(x,  y)i+  N(x,  y)j;  that  is,  assume  F  is  independent 
of  z  and  that  its  k-component  is  identically  zero.) 

1 1 .  Let  S  be  the  surface  defined  by  y  =  10  —  x2  —  z2  with 
y  >  1,  oriented  with  rightward-pointing  normal.  Let 

F  =  (2xyz  +  5z)  i  +  ex  cos  yz  j  +  x2y  k. 

Determine 


V  x  F  •  dS. 


(Hint:  You  will  need  an  indirect  approach.) 

12.  Let  S  be  the  surface  defined  as  z  =  4  —  4x2  —  y2 
with  z  >  0  and  oriented  by  a  normal  with  nonnegative 
k-component.  Let  F(x,  y,  z)  =  x3i  +  ey  j  +  zexyk.. 
Find  UjVxF'  dS.  (Hint:  Argue  that  you  can  inte- 
grate over  a  different  surface.) 

13.  (a)  Show  that  the  path  x(f)  =  (cos?,  sin?,  sin  It)  lies 

on  the  surface  z  =  2xy. 

(b)  Evaluate 


(y  +  cos  x)  dx  +  (sin  y  +  z  )  dy  +  x  dz, 


where  C  is  the  closed  curve  parametrized  and  ori- 
ented by  the  path  x  in  part  (a). 


Chapter  7  I  Surface  Integrals  and  Vector  Analysis 


14.  Let  S  consist  of  the  four  sides  and  the  bottom  face  of 
the  cube  with  vertices  (±a,  ±a,  ±a).  Orient  S  with 
outward-pointing  normals.  Evaluate  //sVxF-JS, 
where  F  =  x2yz3  i  +  x2y  j  +  xex  sin  yz  k. 

1 5.  Use  Stokes's  theorem  to  find  the  work  done  by  the  vec- 
tor field  F  =  (x yz  —  ex)  i  —  xyz  j  +  (x2yz  +  sin  z)  k 
on  a  particle  that  moves  along  the  line  segments  from 
(0,  0,0),  then  to  (1,  1,  1),  then  to  (0,  0,  2),  then  back  to 
(0,  0,  0). 

16.  Let  C  be  a  simple,  closed  curve  that  lies  in  the  plane 
2x  —  3y  +  5z  =  17.  Show  that  the  line  integral 


(3  cos x  +  z)  dx  +  (5x  —  ey)dy  —  3 y  dz 


depends  only  on  the  area  enclosed  by  C  and  its  ori- 
entation, not  on  its  particular  shape  or  location  in  the 
plane. 

1 7.  Use  Gauss's  theorem  to  find  the  volume  of  the  solid  re- 


gion bounded  by  the  paraboloids : 
z  =  3x2+3y2-16. 


y  and 


18.  Let  S  be  defined  by  z  =  el~x  ~y  ,  z  >  1,  oriented 
by  upward  normal,  and  let  F  =  x  i  +  y  j  +  (2  —  2z)  k. 
Use  Gauss's  theorem  to  calculate 


F-dS. 


19.  Give  a  proof  of  Stokes's  theorem  for  smooth, 
parametrized  surfaces  S  =  X(D),  where  X:  D  c 
R2  — >  R3.  To  make  the  proof  easier,  assume  that  X 
is  of  class  C2  and  that  it  is  one-one  on  D  (in  which 
case  3S  =  X(3D)). 

20.  Use  Gauss's  theorem  to  evaluate 


F  •  dS, 


where  F  =  zex  i  +  3y  j  +  (2  —  yz  )  k  and  S  is  the 
union  of  the  five  "upper"  faces  of  the  unit  cube 
[0,  1]  x  [0,  1]  x  [0,  1].  That  is,  the  z  =  0  face  is  not 
part  of  S.  (Hint:  Note  that  S  is  not  closed,  so  to  apply 
Gauss's  theorem  you  will  have  to  close  it  up.) 

21.  In  this  problem,  let  f(x,  y,  z)  be  a  scalar- valued  func- 
tion of  class  C1,  and  let  D  be  a  region  in  space  to 
which  Gauss's  theorem  applies.  Let  n  =  (m,  «3) 
be  the  outward  unit  normal  vector  to  S  =  3D. 

(a)  If  a  is  any  constant  vector  and  F  =  /a,  show  that 
V-F  =  V/-a. 

(b)  Use    part    (a)    with    a  =  i    to    show  that 
3/ 


part 

fn\ dS ■ 


III. 


dV.  Also  obtain  similar 

S  JJJD°* 

results  by  letting  a  equal  j  and  k. 


(c)  Define  a  vector  quantity  (ft)  f  dS 


fndS 


fndSby 


fmds,  (fh  fn2ds,  <th  fmds 

S  w  S  vv  s 


With  notations  and  definitions  as  above,  show  that 


if-ds=ffL 


VfdV. 


(Note  that  the  right  side  is  a  triple  integral  of  a 
vector-valued  expression,  so  it  is  also  computed 
by  integrating  each  scalar  component  function.) 

22.  Given  a  liquid  with  constant  density  S,  introduce  co- 
ordinates so  that  the  (flat)  surface  of  the  liquid  is  the 
xy-plane  and  the  z-coordinate  measures  the  depth  of 
the  liquid  from  the  surface.  (That  is,  the  positive  z-axis 
points  down  into  the  liquid. )  Then  the  pressure  p  inside 
the  fluid  due  to  gravity  is  given  by  p(x,  y,  z)  =  Sgz, 
where  g  is  acceleration  due  to  gravity.  Suppose  that  a 
solid  object  is  immersed  in  the  liquid.  If  the  object  fills 
out  a  region  D  in  space,  then  the  total  buoyant  force 
on  the  solid  is  the  total  liquid  pressure  on  the  boundary 
surface  S  =  3D  and  is  given  by 


B 


pndS, 


23. 


where  n  is  the  outward  unit  normal  to  S.  (The  negative 
sign  arises  because  the  pressure  causes  a  force  point- 
ing inward  on  the  object.)  Use  the  previous  exercise  to 
demonstrate  Archimedes'  principle:  The  magnitude  of 
the  total  buoyant  force  on  an  object  equals  the  weight 
of  the  liquid  displaced. 

Write  a  careful  proof  of  the  three-dimensional  case  of 
Theorem  3.5  of  Chapter  6:  If  F  is  a  vector  field  of  class 
C1  whose  domain  is  a  simply-connected  region  R  in 
R3,  then  F  =  V/  for  some  (scalar- valued)  function  / 
on  R  if  and  only  if  V  x  F  =  0  at  all  points  of  R. 

24.  Let  Sr  denote  the  sphere  of  radius  r  with  center  at  the 
origin,  oriented  with  outward  normal.  Suppose  F  is  of 
class  C 1  on  all  of  R3  and  is  such  that 


F-  dS  =  ar  +  b 


for  some  fixed  constants  a  and  b. 
(a)  Compute 


///„ 


V-FdV, 


where  D  =  {(x,  y,  z)  |  25  <  x2  +  y2  +  z2  <  49}. 
(Your  answer  should  be  in  terms  of  a  and  b.) 

(b)  Suppose,  in  the  situation  just  described,  that  F  = 
V  x  G  for  some  vector  field  G  of  class  C1 .  What 
conditions  does  this  place  on  the  constants  a  and  i>? 


7.3  |  Exercises  507 


25.  Letnfx,  y,  z)  be  a  unit  normal  to  a  surface  S.  The  direc- 
tional derivative  of  a  differentiable  function  f(x,  y,  z) 
in  the  direction  of  n  is  called  a  normal  derivative  of 
/,  denoted  df/dn.  From  Theorem  6.2  of  Chapter  2, 
we  have 

9/ 


V/-n. 


(a)  Let  5  denote  the  portion  of  the  sphere  x2  +  y2  + 
z2  =  a2  in  the  first  octant  (i.e.,  where  x  >  0, 
y  >  0,  z  >  0),  oriented  by  the  unit  normal  that 
points  away  from  the  origin.  Let  f(x,  y,  z)  = 
In  (x2  +  y2  +  z2).  Evaluate 


9/ 
dn 


dS. 


(b)  Let  D  denote  the  piece  of  the  solid  ball  x2  +  y2  + 
z2  <  a2  in  the  first  octant;  that  is, 


D 


[(x,y,z)  \x2  +  y2  +  z2  <a2, 
x  >  0,  y  >  0,  z  >  0}. 


Compute  fffD  V  •  (Vf)dV,  where  /  is  as  in 
part  (a). 

(c)  Apply  Gauss's  theorem  to  the  integral  in  part  (b), 
and  reconcile  your  result  with  your  answer  in 
part  (a). 

26.  Suppose  that  /  is  such  that  for  any  closed,  oriented 
surface  S, 

3/ 

—  dS  =  0. 
dn 

(See  Exercise  25  for  the  definition  of  the  normal  deriva- 
tive df/dn.)  Show  that  then 

d2f      d2f  d2f 

— -  H          H          =  0 

dx2       dy2  dz2 

(i.e.,  that  /  is  harmonic). 

27.  Following  Proposition  3.4,  show  that 

divF(/>)  =  lim  -  <fj)F-dS, 

v^o  V  JTs 

where  S  is  a  piecewise  smooth,  orientable,  closed  sur- 
face 5  enclosing  a  region  D  of  volume  V .  (Take  S  to 
be  oriented  by  outward  normal.)  The  limiting  process 
should  be  assumed  to  be  such  that  D  shrinks  down  to 
the  point  P. 

28.  Use  the  result  of  Exercise  27  to  establish  the  formula 
for  the  divergence  of  a  C1  vector  field 

F  =  Fx(x,  y,  z)i+  F2(x,  y,  z)j  +  F3(x,  y,  z)k. 

That  is,  show  that 

dF\      dFi  dF-i 
divF=  — !-  +  —  +  — 

dx        dy  dz 


in  the  following  manner:  Let  P  have  coordinates 
(*o,  yo,  zo)  and  consider  the  (small)  cube  S,  of  edge 
length  a,  centered  at  P  with  faces  parallel  to  the  coor- 
dinate planes.  Note  that  the  volume  V  enclosed  by  S 
is  ai.  It  will  help  to  recall  that  if  f(x,  y,  z)  is  differen- 
tiable, then 


dx 


lim 


f(x  +  Ax,y,z)-  f{x,y,z) 
Ax 


=  lim 


f(x  +  ^,y,z)-f{x 


Ax 
2  ' 


y.  z) 


Ax 


29.  In  this  problem,  you  will  use  the  result  of  Exercise  27  to 
find  an  expression  for  V  •  F  in  cylindrical  coordinates. 
(See  Theorem  4.5  of  Chapter  3.)  Begin  by  writing 

F  =  Fr  e,  +  Fe  ee  +  Fz  ez , 

where  Fr(r,  9,  z),  Fg{r,  6,  z),  and  Fz(r,  9,  z)  denote  the 
components  of  F  in  the  e,--,  eg-,  and  e; -directions 
(respectively).  Let  P  have  cylindrical  coordinates 
(r,9,z).  Consider  the  small  "cylindrical  coordinate 
cuboid"  S  shown  in  Figure  7.45.  The  pairs  of  oppo- 
site faces  correspond  to  values 

r  -  Ar/2  and  r  +  Ar/2; 
9-A9/2  and  9  +  A9 /2; 
z  -  Az/2     and     z  +  Az/2. 

Note  that  the  volume  of  the  cuboid  is  approximately 
r  A9  Ar  Az. 

(a)  Approximate  §s  F  •  dS  (where  S  is  oriented  by  out- 
ward unit  normal)  by  noting  that  each  face  of  S  is 
roughly  flat  with  an  "obvious"  unit  normal  vector 
and  that  F  is  approximately  constant  on  each  face. 

(b)  Use  your  answer  in  part  (a)  to  calculate  the  diver- 
gence in  cylindrical  coordinates  as 


 1  (rFr)  H  . 

dz       r  dr  r  d9 


divF 

(This  agrees  with  formula  (4)  of  §3.4.) 


Figure  7.45  The  cylindrical 
coordinate  cuboid  of 
Exercise  29. 


Chapter  7  I  Surface  Integrals  and  Vector  Analysis 

30.  Use  the  ideas  of  Exercises  27  and  29  to  calculate  the 
divergence  in  spherical  coordinates.  (See  Theorem  4.6 
of  Chapter  3.)  You  will  want  to  make  use  of  the  small 
"spherical  coordinate  cuboid"  S  shown  in  Figure  7.46. 


31, 


Figure  7.46  The  spherical  coordinate 
cuboid  of  Exercise  30.  The  volume  of 
the  cuboid  is  approximately 
p2  simp  AO  A<p  Ap. 
The  pairs  of  opposite  faces  of  S  correspond  to  values 


p  - 

Ap/2 

and 

P  +  Ap/2; 

<p  - 

Acp/2 

and 

<P  +  A<p/2; 

e  - 

AO/2 

and 

0  +  Ad  12. 

Let  F  be  a  vector  field  of  class  C  in  a  neighborhood 
of  the  point  P  in  R3 ,  and  let  n  be  a  unit  vector  drawn 
with  its  tail  at  P .  Let  C  be  a  simple,  closed  curve  such 
that  there  is  an  orientable  surface  S  bounded  by  C  that 
contains  P  and  such  that  n  is  normal  to  S  at  P.  Orient 
S  by  using  n,  and  orient  C  consistently  with  S.  Follow- 
ing Proposition  3.5,  show  that,  if  A  denotes  the  area  of 
S,  then 


n-curlF(P) 


1 

lim  — 

A^O  A 


F  •  ds. 


Here  the  limiting  process  is  assumed  to  be  such  that  C 
shrinks  down  to  the  point  P.  (See  Figure  7.47.) 


(a)  Find  the  ez-component  of  curlF  by  considering 
the  planar  path  shown  in  Figure  7.48.  The  pairs  of 
opposite  "edges"  of  the  approximately  rectangular 


Figure  7.48  The  path  C  of 
Exercise  32(a). 

path  C  correspond  to  the  values 

r  —  Ar/2    and    r  +  Ar/2, 

and 

0-A0/2    and    6  +  A6/2 

(all  with  constant  z-coordinate).  Note  that  the  area 
enclosed  by  C  is  approximately  r  AO  Ar.  Approx- 
imate the  line  integral  fc  F  •  ds  by  using  the  fact 
that,  for  small  AO  and  Ar,  each  edge  of  C  is 
roughly  straight.  Show  that 

1  dFr  13 


•  curl  F  : 


+  (rFg). 

r  d0       r  or 


(b)  Use  the  path  in  Figure  7.49  to  show  that 
1  dFz  dFe 
r~d0  dz' 


er  •  curl  F  : 


Figure  7.47  Figure  for 
Exercise  3 1 . 

32.  In  this  problem,  you  will  use  the  result  of  Exercise  3 1 
to  determine  an  expression  for  curl  F  in  cylindrical  co- 
ordinates. Begin  by  writing 


Figure  7.49  The  path  C  of 
Exercise  32(b). 


(c)  Use  the  path  in  Figure  7.50  to  show  that 

dFr  dFz 
dz  dr 

Combine  this  with  the  results  of  parts  (a)  and  (b) 
to  obtain 


eg  •  curl  F  = 


curlF  = 


1 


er 

ree  ez 

d/dr 

d/90  d/dz 

Fr 

rFg  Fz 

F  =  Fr  er  +  Fe  e9  +  Fz  ez 


(See  Theorem  4.5  of  Chapter  3.) 


7.3  |  Exercises 


Az 


Ay- 


Figure  7.50  The  path  C  of 
Exercise  32(c). 

33.  In  this  problem,  you  will  determine  an  expression  for 
curl  F  in  spherical  coordinates.  Let  F  be  a  vector  field 
of  class  C1,  and  write 

F  =  Fpep  +  Fve(p  +  F0ee. 

Show  that 

1 


ep  •  curl  F 


p  smcp 


9  9F„ 
— (sinpFfl)-  — — 
dip  86 


and 


e9  •  curl  F 


eg  •  curl  F 


1  dFp  9 
sinip  90  9p 


dp 


(pF<p) " 


3f„ 
9^ 


by  using  Exercise  3 1  and  the  three  paths  shown  in  Fig- 
ure 7.51.  Conclude  that 


curl  F 


1 


pL  sin  <p 


p  sin  <p  eg 

9/9p 

d/d(p 

d/de 

Fp 

pF<p 

p  sin  (p  Fe 

(See  Theorem  4.6  of  Chapter  3.) 

34.  Of  the  six  planar  vector  fields  shown  in  Figure  7.52, 
four  have  zero  divergence  in  the  regions  indicated  and 
three  have  zero  curl.  By  considering  appropriate  in- 
tegrals and  using  the  results  of  Exercises  27  and  3 1 , 
categorize  each  vector  field. 


Figure  7.51  The  paths  of  Exercise  33. 


;•!»!»;• 


\  1 1' i , 


(b) 


(d) 


Figure  7.52  Figures  for  Exercise  34. 


Chapter  7  i  Surface  Integrals  and  Vector  Analysis 


7.4  Further  Vector  Analysis;  Maxwell's 
Equations 

In  this  section,  we  use  Gauss's  theorem  and  Stokes's  theorem  first  to  prove  some 
abstract  results  in  vector  analysis  and  then  to  study  Maxwell's  equations  of  elec- 
tricity and  magnetism. 


Green's  Formulas   

Our  purpose  in  §7.4  is  to  establish  a  few  fundamental  results  of  vector  analysis. 
Throughout  the  discussion  all  scalar  and  vector  fields  are  defined  on  subsets  of  R3 . 
The  following  pair  of  results  is  established  readily: 


THEOREM  4.1  (Green's  first  and  second  formulas)  Let  /  and  g  be  scalar 
fields  of  class  C2,  and  let  D  be  a  solid  region  in  space,  bounded  by  a  piecewise 
smooth  surface  S  =  dD,  oriented  as  in  Gauss's  theorem.  Then  we  have 

Green's  first  formula: 

/  / fDVf-V8dV  +  /  /  fDfV2gdV  =  §sfV8'dS- 
Green's  second  formula: 


III 


(fV2g-gV2f)dV  =  (fb(fVg-gVf) -dS. 

D  MS 


PROOF  The  first  formula  follows  from  Gauss's  theorem  applied  to  the  vector 
field  F  =  /Vg.  (We  leave  the  details  to  you.)  The  second  formula  follows  from 
writing  the  first  formula  twice: 

/  fLvf'v§dv+ /  /  Lfv28dv= §sfV8'ds'  (1) 
/ ILV8'vfdv+ 1 1 L8V2fdv=§s8Vf'ds'  (2) 

Now,  subtract  equation  (2)  from  equation  (1).  ■ 
The  third  of  Green's  formulas  requires  considerably  more  work  to  prove. 


THEOREM  4.2  (Green's  third  formula)  Assume  /  is  a  function  of  class  C2. 
Then,  for  dD  =  S  oriented  as  in  Gauss's  theorem  and  points  r  in  the  interior 
ofD, 

1    f  f  f  V2  / (x) 

4u K\  JW  V hi-  —  xii y   mi-  —  xii 7 

In  this  formula,  d  V  denotes  integration  with  respect  to  the  variables  in  x  = 
(x,  y,  z)  (i.e.,  r  =  (ri,  r2,  r3)  is  fixed  in  the  integration),  and  the  symbol  V  means 
Vx,  differentiation  with  respect  to  x,  y,  and  z. 


A  proof  of  Theorem  4.2  appears  in  the  addendum  to  this  section. 


7.4  |  Further  Vector  Analysis;  Maxwell's  Equations 


An  Inversion  Formula  for  the  Laplacian  

Green's  third  formula  is  a  type  of  inversion  formula — a  formula  that  enables 
us  to  recover  the  values  of  a  function  /  by  knowing  certain  integrals  involving 
its  gradient  and  Laplacian.  Green's  third  formula  is  mainly  of  technical  interest 
in  proving  further  results.  We  use  it  here  to  obtain  an  inversion  formula  for  the 
Laplacian  operator. 

We  begin  by  applying  the  Laplacian  V2  to  Green's  third  formula: 


Vr2/(r)  =  Vr2 


1  [[[  Yl 

in  J  J  Jd  llr- 


xJ  —  dV 

An 


(3) 


The  trick  is  to  move  V2  inside  the  surface  integral,  which  is  justified  since  x^r 
when  x  varies  over  S: 


V^t^-/(x)V*fe»»-JS 


s      \l|r-x||  \||r-x" 
By  direct  calculation,  V2(l/||r  —  x||)  =  0  for  x  ^  r,  so 

'  Vllr-il 
Similarly,  since  /(x)  does  not  involve  r, 

Vr2  (7(x)Vx  f^-*— -  H  =  /(x)Vr2Vx  / 


=  /(x)VxVr2 


|r-x| 
1 


=  0. 

Therefore,  the  Laplacian  of  the  original  surface  integral  is  0.  We  may  conclude 
from  equation  (3)  that,  for  r  in  the  interior  of  D, 

4jt    r  JJJd  ||r-x|| 
and  we  have  shown  the  following: 

THEOREM  4.3  If  <p  =  V2/  for  some  function  /  of  class  C2,  then  for  r  in  the 
interior  of  D, 


-Lv2  [[[  Jt 

to  rJJJD\\r 


Theorem  4.3  provides  an  inversion  formula  for  the  Laplacian  in  the  following 
sense:  If  V2/  =  <p,  then 

where  g  is  any  harmonic  function  (i.e.,  g  is  such  that  V2g  =  0).  That  is,  if  the 
Laplacian  of  /  is  known,  we  can  recover  the  function  /  itself,  up  to  addition  of 
a  harmonic  function. 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Figure  7.53  The  sphere  Sj, 
of  radius  b,  centered  at  the 
origin,  together  with  an 
outward  unit  normal  vector. 


Finally,  it  can  be  shown  that  the  result  of  Theorem  4.3  holds  under  consider- 
ably less  stringent  hypotheses  than  having  <p  be  the  Laplacian  of  another  function. 1 

Maxwell's  Equations   


Maxwell's  equations  are  fundamental  results  that  govern  the  behavior  of — and 
interactions  between — electric  and  magnetic  fields.  We  see  how  Maxwell's  equa- 
tions arise  from  a  few  simple  physical  principles  coupled  with  the  vector  analysis 
discussed  previously. 

Gauss's  law  for  electric  fields  If  E  is  an  electric  field  then  the  flux  of  E 
across  a  closed  surface  S  is 


Flux  of  E  =  (fl)E-dS. 


(4) 


(5) 


Applying  Gauss's  theorem  to  formula  (4),  we  find  that 

Flux  of  E  =  jjj  V-EdV, 

where  D  is  the  region  enclosed  by  S. 

If  the  electric  field  E  is  determined  by  a  single  point  charge  of  q  coulombs 
located  at  the  origin,  then  E  is  given  by 

q  x 


E(x) 


(6) 


4jt€0  ||X|| 

where  x  =  x  i  +  y  j  +  z  k.  In  mks  units,  E  is  measured  in  volts/meter.  The  constant 
€q  is  known  as  the  permittivity  of  free  space;  its  value  (in  mks  units)  is 
8.854  x  10-12  coulomb2 /newton-meter2. 

For  the  electric  field  in  equation  (6),  we  can  readily  verify  that  V  •  E  =  0  wherever 
E  is  defined.  From  formulas  (4)  and  (5),  if  S  is  any  surface  that  does  not  enclose 
the  origin,  then  the  flux  of  E  across  S  is  zero. 

But  now  a  question  arises:  How  do  we  calculate  the  flux  of  the  electric  field 
in  equation  (6)  across  surfaces  that  do  enclose  the  origin?  The  trick  is  to  find 
an  appropriate  way  to  exclude  the  origin  from  consideration.  To  that  end  first 
suppose  that  Sb  denotes  a  sphere  of  radius  b  centered  at  the  origin  (i.e.,  Si,  has 
equation  x2  +  y2  +  z2  =  b2).  Then  the  outward  unit  normal  to  Si,  is 

xi  +  yj  +  zk  1 


=  -  x. 


(See  Figure  7.53.)  From  equation  (6), 


On  Sb,  we  have  ||x|| 


E-dS  = 

Si, 

b,  so  that 


E-dS  = 


Si, 


4jre0 
x 

b~l' 


-xdS. 


Si, 


Si, 


-dS=  —!— 
b  4tt€0 


Si, 


ixr 
b4 


dS 


4jTe0b2 


47T€0b2 


_  1 

¥      ~  Aneob2 


(surface  area  of  Sb) 


dS 


Si, 


(4jvb2) 


'  See,  for  example,  0.  D.  Kellogg,  Foundations  of  Potential  Theory  (Springer,  1928;  reprinted  by  Dover 
Publications,  1954),  p.  220. 


7.4  |  Further  Vector  Analysis;  Maxwell's  Equations 


Figure  7.54  The  solid  region  D  is 
the  region  inside  S  and  outside  S/,. 


Now,  suppose  S  is  any  surface  enclosing  the  origin.  Let  Si,  be  a  small  sphere 
centered  at  the  origin  and  contained  inside  S.  Let  D  be  the  solid  region  in  R3 
between  Sb  and  5.  (See  Figure  7.54.)  Note  that  V  •  E  =  0  throughout  D,  since 
D  does  not  contain  the  origin.  (E  is  still  denned  as  in  equation  (6).)  Orienting 
dD  =  S  U  Sb  with  normals  that  point  away  from  D,  we  obtain 


(ffiE-dS-  £  E-dS=  £   E  •  dS  =  f  f  f 

MS  JJsi,  JJdD  J  J  J  D 


V-EdV  =  0, 


using  Gauss's  theorem.  We  conclude  that 


E  •  dS 


q_ 


(7) 


for  any  surface  that  encloses  the  origin.  By  modifying  equation  (6)  for  E  appro- 
priately, we  can  show  that  formula  (7)  also  holds  for  any  closed  surface  containing 
a  single  point  charge  of  q  coulombs  located  anywhere.  (Note:  The  arguments  just 
given  hold  for  any  inverse  square  law  vector  field  F(x)  =  &x/||x||3,  where  k  is  a 
constant.  See  Exercise  1 3  in  this  section  for  details.) 

We  can  adapt  the  arguments  just  given  to  accommodate  the  case  of  n  discrete 
point  charges.  For  i  =  1 , . . . ,  n,  suppose  a  point  charge  of  coulombs  is  located 
at  position  r, .  The  electric  field  E  for  this  configuration  is 


E(x)  = 


1 


47T60 


13  • 


(8) 


For  E  as  given  in  equation  (8),  we  can  calculate  that  V  •  E  =  0,  except  at  x  =  r, . 
If  S  is  any  closed  piecewise  smooth,  outwardly-oriented  surface  containing  the 
charges,  then  we  may  use  Gauss's  theorem  to  find  the  flux  of  E  across  S  by  taking 
n  small  spheres  Si,  S%, . . . ,  Sn,  each  enclosing  a  single  point  charge.  (See  Fig- 
ure 7.55.)  If  D  is  the  region  inside  S  but  outside  all  the  spheres,  we  have,  by 
choosing  appropriate  orientations  and  using  Gauss's  theorem, 

dffE-JS-V<jffE-JS=(jff  E-dS=  f  f  f  V-EdV  =  0, 

MS  j_i  MS;  JJdD  J  J  JD 

since  V  •  E  =  0  on  D.  Hence, 


E-dS  =  ^2<jj)  E  •  dS  =  —  1i 


1 

=  —  (total  charge  enclosed  by  S). 


(9) 


Figure  7.55  D  is  the  solid  region  inside  the 

surface  S  and  outside  the  small  spheres 

Si ,  S2  Sn,  each  enclosing  a  point  charge. 


514       Chapter  7  i  Surface  Integrals  and  Vector  Analysis 

To  establish  Gauss's  law,  consider  the  case  not  of  an  electric  field  determined 
by  discrete  point  charges,  but  rather  of  one  determined  by  a  continuous  charge 
distribution  given  by  a  charge  density  p(x).  The  total  charge  over  a  region  D  in 
space  is 

so  that,  in  place  of  formula  (8),  we  have  an  electric  field, 

E(r)=  —  f  f  f  p(x)  T~XdV.  (10) 
IntJJjD  l|r-x||3 

In  equation  (10),  the  integration  occurs  with  respect  to  the  variables  in  x.  (Note: 
It  is  not  at  all  obvious  that  the  integral  used  to  define  E(r)  converges  at  points 
r  e  D,  where  p(r)  ^  0,  because  at  such  points  the  triple  integral  in  equation  (10) 
is  improper.  See  Exercise  20  in  this  section  for  an  indication  of  how  to  deal  with 
this  issue.) 

The  integral  form  of  Gauss's  law,  analogous  to  that  of  formula  (9),  is 

E  E-dS  =  —  [  [  [  pdV,  (11) 
JJs  eo  J  J  Jd 

where  5  =  3D.  If  we  apply  Gauss's  theorem  to  the  left  side  of  formula  (1 1),  we 
find  that 

[  [  [  V  -EdV  =  —  1 1 1  pdV. 
J  J  Jd  to  J  J  Jd 

Since  the  region  D  is  arbitrary,  it  may  be  "shrunk  to  a  point."  From  this,  we 
conclude  that 


p 

V-E  =  — . 

(12) 

Equation  (12)  is  the  differential  form  of  Gauss's  law. 

Magnetic  fields  A  moving  charged  particle  generates  a  magnetic  field.  To  be 
specific,  if  a  point  charge  of  q  coulombs  is  at  position  ro  and  is  moving  with 
velocity  v,  then  the  magnetic  field  it  induces  is 

In  mks  units,  B  is  measured  in  teslas.  The  constant  /xo  is  known  as  the  perme- 
ability of  free  space;  in  mks  units 

Ho  =  Ait  x  10~7  N/amp2  «  1.257  x  10~6  N/amp2. 

In  the  case  of  a  magnetic  field  that  arises  from  a  continuous,  charged  medium 
(such  as  electric  current  moving  through  a  wire),  rather  than  from  a  single  moving 
charge,  we  replace  q  by  a  suitable  charge  density  function  p  and  the  velocity  of 
a  single  particle  by  the  velocity  vector  field  v  of  the  charges.  Then  we  define  the 
current  density  field  J  by 


J(x)  =  p(x)v(x). 


(14) 


7.4  |  Further  Vector  Analysis;  Maxwell's  Equations 


Figure  7.56  The 

total  current  / 
across  S  is  the  flux 
of  the  current 
density  J  across  S. 


In  place  of  formula  (13),  we  use  the  following  definition  for  the  magnetic  field 
resulting  from  moving  charges  in  a  region  D  in  space: 


B(r)  =  s////<I)vW* 
///DJWx 


dV 


Mo 


r-x\ 
r  —  x 


dV. 


(15) 


In  equation  (15),  as  in  equation  (10),  the  integration  is  with  respect  to  the  vari- 
ables constituting  x.  As  in  equation  (10),  it  is  not  obvious  that  the  integrals  in 
equation  (15)  are  convergent  if  r  e  D,  but,  in  fact,  they  are.  (See  Exercise  21  in 
this  section.) 

Before  continuing  our  calculations,  we  comment  further  regarding  the  current 
density  field  J.  The  vector  field  J  at  a  point  is  such  that  its  magnitude  is  the  current 
per  unit  area  at  that  point,  and  its  direction  is  that  of  the  current  flow.  It  is  not 
hard  to  see  then  that  the  total  current  /  across  an  oriented  surface  S  is  given  by 
the  flux  of  J;  that  is, 


/ 


"//. 


J-dS. 


(16) 


(See  Figure  7.56.) 

Returning  to  the  magnetic  field  B  in  equation  (15),  we  show  that  it  can  be 
identified  as  the  curl  of  another  vector  field  A  (to  be  determined).  First,  by  direct 
calculation, 

1 


|r-x| 

Therefore,  equation  (15)  becomes 


B(r) 


Mo 
An 


Iff. 


J(x)  x  Vr 


dV. 


(17) 


We  rewrite  equation  (17)  using  the  following  standard  (and  readily  verified)  iden- 
tity, where  /  is  a  scalar  field  and  F  a  vector  field  (both  of  class  C2): 


V  x  (/F)  =  (V  x  F)/  -  F  x  V/. 
Formula  (18)  is  equivalent  to 

FxV/=(VxF)/-Vx  (/F). 

Therefore, 

1     \      .  .1 


(18) 


J(x)  x  Vr 


=  (Vr  x  J(x)) 


Vr  x 


J(x) 


V,  x 


J(x) 


J(x) 


since  J(x)  is  independent  of  r.  Hence, 

AttJJJd         ||r-x||  4;r         J  J  Jd 

as  r  does  not  contain  any  of  the  variables  of  integration.  Consequently, 

B(r)  =  V  x  A(r), 


dV, 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


C 


Figure  7.57  The 

closed  loop  C  is 
oriented  so  that  it 
has  a  right-hand 
relationship  with 
the  direction  of 
current  flow  it 
encloses. 


where 


A(r) 


IM) 


1 1 Id  II  r 


J(x) 


dV. 


(19) 


4tt  J  J  J d  l|r-x| 

Thus,  V  •  B  =  V  •  ( V  x  A)  and  so,  by  Theorem  4.4  in  Chapter  3,  we  conclude  that 


V-B  =  0. 


(20) 


The  intuitive  content  of  equation  (20)  is  often  expressed  by  saying  that  "magnetic 
monopoles"  do  not  exist. 

The  vector  field  A  in  equation  ( 1 9)  furnishes  an  example  of  a  vector  potential 
of  the  field  B.  (See  Exercises  33-38  in  the  Miscellaneous  Exercises  for  more  about 
vector  potentials.) 

Ampere's  circuital  law  If  C  is  a  closed  loop  enclosing  a  current  /,  then 
Ampere's  law  says  that,  up  to  a  constant,  the  current  through  the  loop  is  equal  to 
the  circulation  of  the  magnetic  field  around  C.  To  be  precise, 


B  •  ds  =  fiQl. 


(21) 


In  equation  (2 1 ),  we  assume  that  C  is  oriented  so  that  C  and  /  are  related  by  a 
right-hand  rule,  that  is,  that  they  are  related  in  the  same  way  that  the  orientation  of 
C  and  the  normal  to  any  surface  S  that  C  bounds  are  related  in  Stokes's  theorem. 
(See  Figure  7.57.) 

From  equation  (16)  for  the  total  current,  equation  (21)  may  be  rewritten  as 

j>  B-ds  =  /i,0  J  j  J-dS, 

where  S  is  any  (piecewise  smooth,  oriented)  surface  bounded  by  C.  Applying 
Stokes's  theorem  to  the  line  integral,  we  obtain 


J  J  V  x  B  •  dS  =  fi0  J  J  J  •  dS. 
Since  the  loop  C  and  surface  S  are  arbitrary,  we  conclude  that 


V  x  B  =  (IqJ. 


(22) 


Equation  (22)  is  the  differential  form  of  Ampere's  law  in  the  static  case 

(i.e.,  in  the  case  where  B  and  E  are  constant  in  time).  In  the  event  that  the  magnetic 
and  electric  fields  are  time  varying,  we  need  to  make  some  modifications.  The  so- 
called  equation  of  continuity,  established  in  Exercise  5  in  this  section,  states  that 

dp 


V-J  = 


(23) 


The  difficulty  is  that  if  V  x  B  =  /zoJ  as  in  equation  (22),  then  equation  (23) 
implies  that 

dp 


V  •  (V  x  B)  =  V  •  QioS) 


-Mo 


dt 


7.4  |  Further  Vector  Analysis;  Maxwell's  Equations 


However,  assuming  B  is  of  class  C2,  we  must  have  V  •  (V  x  B)  =  0,  even  in  the 
case  where  p  is  not  constant  in  time. 

The  simplest  solution  to  this  difficulty  is  to  modify  equation  (22)  by  adding 
an  extra  term.  From  Gauss's  law,  equation  (12),  we  must  have 

dp  9E 
dt  dt 

Thus,  if  we  replace  J  by  J  +  eo(dE/dt)  in  equation  (22),  then  we  can  verify  that 
V  •  (V  x  B)  =  0.  (See  Exercise  16  in  this  section.)  Hence,  Ampere's  law  can  be 
generalized  as 


V  X  B  =  /X0J  +  tlQ€o 


9E 

dt' 


(24) 


Figure  7.58  The 

rate  of  change  of 
magnetic  flux  across 
S  determines  the 
electromotive  force 
around  the 
boundary  C. 


The  term  e0(9E/9/)in  equation  (24),  known  as  the  displacement  current  density, 
was  first  postulated  by  James  Clerk  Maxwell  in  order  to  generalize  Ampere's  law 
to  the  nonstatic  case.  (In  this  context,  the  original  current  density  field  J  is  known 
as  the  conduction  current  density.) 

Equation  (24)  is  not  the  only  possible  generalization  of  equation  (22),  but 
it  is  the  simplest  one  and  is  consistent  with  observation.  See  Exercise  17  in  this 
section  for  other  ways  to  generalize  equation  (22)  to  the  nonstatic  case. 

Faraday's  law  of  induction  Michael  Faraday  observed  empirically  that  the 
change  in  magnetic  flux  across  a  surface  S  equals  the  electromotive  force  around 
the  boundary  C  of  the  surface.  This  relation  can  be  written  as 


d1> 
dt 


E-Js, 


(25) 


where  4>(f)  =  ffs  B  •  dS,  and  C  and  S  are  oriented  consistently.  (See  Figure  7.58.) 
If  we  apply  Stokes's  theorem  to  the  line  integral,  we  find  that 


V  x  E  •  dS. 


Since 


we  have 


d<S> 
dt 


=  —  f  f  B  dS=  f  f  —  -dS, 
dt  J  Js  J  Js  dt 


V  x  E-dS. 


is  ^  J  Js 

Because  equation  (26)  holds  for  arbitrary  surfaces,  we  conclude  that 


(26) 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Summary  Equations  (12),  (20),  (24),  and  (27)  together  are  known  as  Maxwell's 
equations: 


V-E  = 

P 

(Gauss's  law); 

eo 

V  B  = 

0 

(No  magnetic  monopoles); 

3B 

V  x  E  = 

(Faraday's  law); 

V  x  B  = 

3E 

Mo  J  +  M-oeo  — 
at 

(Ampere's  law). 

Maxwell's  equations  allow  us  to  reconstruct  the  electric  and  magnetic  fields  from 
the  charge  and  current  densities.  They  are  fundamental  to  the  subject  of  electricity 
and  magnetism  and  provide  a  fitting  tribute  to  the  power  of  the  theorems  of  Stokes 
and  Gauss. 


Addendum:  Proof  of  Theorem  4.2   

The  most  obvious  idea  is  to  use  Green's  second  formula  with 


However,  this  function  fails  to  be  continuous  when  x  =  r,  so  Gauss's  theorem 
(and  hence  Green's  formula)  cannot  be  applied  so  readily.  Instead  we  need  to 
examine  the  integrals  more  carefully. 

Throughout  the  discussion  that  follows,  let  Sb  denote  the  sphere  of  radius  b 
centered  at  r.  First,  we  establish  some  subsidiary  results. 


■  Lemma  1   If  h  is  a  continuous  function,  then 

hm$  J^-dS  =  0. 
b^JJSb  Hr-x|| 


Figure  7.59  Ifx 

is  any  point  on  the 
surface  of  the 
sphere  of  radius  b 
centered  at  r,  then 
llr  —  xll  =  b. 


PROOF  The  average  value  of  h  on  Sb  is  defined  to  be 

§Sbh{,)dS  i  rr 

[hU  =  surface  area  of  5,  =  W  % 
(See  Exercise  9  of  the  Miscellaneous  Exercises.)  Thus, 


h(x)dS. 


h{x)dS  =  Anb2[h\ 


As  x  varies  over  the  surface  Sb,  we  have  ||r  —  x||  =  b.  (See  Figure  7.59.)  Hence, 


lim 


h(x) 


dS  =  lim 


1 


b-*0  .JJSb  b 


h(x)dS  =  lim  4nb  \h\w„ 

b->-0 


To  clarify  the  variables  with  respect  to  which  we  differentiate,  let  Vx  denote 
the  del  operator  with  respect  to  x,  y,  and  z,  and  Vr  the  del  operator  with  respect 
to  r  =  (n,  r2,  r3). 


7.4  |  Further  Vector  Analysis;  Maxwell's  Equations 


Figure  7.60  The  region 
D  —  B  denotes  the  solid  D 
with  a  small  ball  centered  at  r 
removed. 


Lemma  2  With  h  and  Sb  as  in  Lemma  1 , 
lim  (th  h(x)Vx  ' 


Si, 


•  dS  =  -4nh(r). 


PROOF  Let  n  =  (x  —  r)/||r  —  x||,  the  normalization  of  x  —  r.  Straightforward 
calculations  yield 


1 


1 


and 
n- Vx 

]|r  -  x 
for  x  on  Si.  Then 

lim  §  /j(x)Vx 


b^O 


Sb 


(x  -  r)  •  (x  -  r) 


dS  =  lim 

b^O 


1 

b5' 


s„ 


h(x)Vx 


n\dS 


lim 

b^O 


Si, 


h(x)dS 


=  l™0-b2(4nb2 

=  -  4nh(r). 


(See  the  proof  of  Lemma  1.) 


Returning  to  the  proof  of  Green's  third  formula,  we  look  at  a  region  to  which 
we  can  apply  Green's  second  formula,  namely,  the  region  D  —  B,  where  B  is  a 
small  ball  of  radius  b  centered  at  r.  (See  Figure  7.60.)  By  Green's  second  formula 
(since  l/||r  —  x||  is  not  singular  on  D  —  B),  we  have 


///. 


dV 


-If 

J  Js-sb 


/(x)Vx 


1 


r  —  x 

Vx/(X) 


■dS. 


(28) 


By  direct  calculation  V^l/Hr  —  x||)  =  0,  so  equation  (28)  becomes 


SSL 


Vx2/(x) 


dV 


s-s„ 


/(x)Vx 


Vx/(x) 


dS. 


(29) 


We  may  evaluate  the  right-hand  side  of  equation  (29)  by  replacing  the  surface 
integral  over  S  —  St  by  separate  integrals  over  S  and  Sb-  Now,  we  take  limits  as 
b  ->•  0.  By  Lemma  1  with  h(x)  =  Vx/(x)  •  n, 


V*/(x) 


dS  = 


Si, 


Vx/OO-n 


dS  ->  0. 


Si, 


By  Lemma  2, 


/(x)Vx 


Si, 


■  dS  -4jr/(r). 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


Since  B  shrinks  to  a  point  as  b  ->  0,  we  see  that  equation  (29)  becomes 


///. 


'  d    i|r-x||  jjs  \  \\\r-x\\;     \\r  ■ 

from  which  Green's  third  formula  follows  immediately. 


7.4  Exercises 


1 .  Prove  Green's  first  formula,  stated  in  Theorem  4. 1 . 

A  function  g(x,  y,  z)  is  said  to  be  harmonic  at  a  point 
(xq,  yo,  Zo)  if  8  is  of  class  C2  and  satisfies  Laplace  s  equation 


V2g 


dx2      dy2  dz2 


0 


on  some  neighborhood  of  (xq,  yo,  zo).  We  say  that  g  is  har- 
monic on  a  closed  region  DCR!  if  it  is  harmonic  at  all  in- 
terior points  of  D  (i.e.,  not  necessarily  on  the  boundary  of 
D).  Exercises  2-4  concern  some  elementary  vector  analysis  of 
harmonic  functions. 

2.  Assume  that  D  is  closed  and  bounded  and  that  3D  is  a 
piecewise  smooth  surface  oriented  by  outward  unit  nor- 
mal field  n.  Let  dg/dn  denote  Vg  •  n.  (The  term  dg/dn 
is  called  the  normal  derivative  of  g.)  Use  Green's  first 
formula  with  f(x,  y,z)  =  1  to  show  that,  if  g  is  har- 
monic on  D,  then 

'  d-lds  =  o. 

dn 

3.  Let  /  be  harmonic  on  a  region  D  that  satisfies  the 
assumptions  of  Exercise  2. 
(a)  Show  that 


///„ 


Vf-VfdV 


3/ 
f—dS. 
6D  dn 


(b)  Suppose  f(x,y,z)  =  0  for  all  (x,  y,  z)  G  dD. 
Use  part  (a)  to  show  that  then  we  must  have 
f(x,  y,  z)  =  0  throughout  all  of  D.  (Hint:  Think 
about  the  sign  of  V/  •  V/.) 

4.  Let  D  be  a  region  that  satisfies  the  assumptions  of 
Exercise  2.  Use  the  result  of  Exercise  3(b)  to  show 
that  if  f\  and  fi  are  harmonic  on  D  and  fi(x,  y,  z)  = 
fi{x,  y,  z)  on  3D,  then,  in  fact,  f\  =  fi  on  all  of  D. 
Thus,  we  see  that  harmonic  functions  are  determined 
by  their  boundary  values  on  a  region.  (Hint:  Consider 

fx  ~  fi.) 

5.  (a)  Suppose  a  fluid  of  density  p(x,  y,  z,  t)  flows  with 

velocity  field  F(x,  y,  z,  t)  in  a  solid  region  W  in 
space  enclosed  by  a  smooth  surface  S.  Use  Gauss's 
theorem  to  show  that,  if  there  are  no  sources  or 
sinks, 


V  •  (pF)  : 


dp 
9? 


This  equation  is  called  the  continuity  equa- 
tion in  fluid  dynamics.  (Hint:  The  triple  integral 
fffw  lr  dV  is  the  rate  of  fluid  flowing  into  W, 
and  the  flux  of  pF  across  S  gives  the  rate  of  fluid 
flowing  out  of  W.) 

(b)  Use  the  argument  in  part  (a)  to  establish  the  equa- 
tion of  continuity  for  current  densities  given  in 
equation  (23): 

dp 
'  dt  ' 


V-J: 


Let  T(x,  y,  z,  t)  denote  the  temperature  at  the  point  (x,  y,  z) 
of  a  solid  object  D  at  time  t.  We  define  the  heat  flux  density  H 
by  H  =  —kVT.  (The  constant  k  is  the  thermal  conductivity. 
Note  that  the  symbol  V  denotes  differentiation  with  respect  to 
x,y,z,  not  with  respect  to  t.)  The  vector  field 'H  represents  the 
velocity  of  heat flow  in  D.  Lt  is  a  fact  from  physics  that  the  total 
heat  contained  in  a  solid  body  D  having  density  p  and  specific 
heat  a  is 


SSL 


apTdV. 


Hence,  the  total  amount  of  heat  leaving  D  per  unit  time  is 

97* 


SSL 


op — dV. 

d  at 


(Here  we  assume  that  a  and  p  do  not  depend  on  1.)  We  also 
know  that  the  heat  flux  may  be  calculated  as 


ss 


n-ds. 


Exercises  6-10  concern  these  notions  of  temperature,  heat,  and 
heat  flux  density. 

6.  Use  Gauss's  theorem  to  derive  the  heat  equation, 

97' 


op 


dt 


kV2T. 


7.  In  Exercise  6,  suppose  that  k  varies  with  the  points  in 
D;  that  is,  k  =  k(x,  y,  z).  Show  that  then  we  have 


op 


dT 
97 


kVzT  +  Vk-VT. 


8.  In  the  heat  equation  of  Exercise  6,  suppose  that  o,  p, 
and  k  are  all  constant  and  the  temperature  T  of  the 
solid  D  does  not  vary  with  time.  Show  that  then  T 
must  be  harmonic,  that  is,  that  V2J  =  0  at  all  points 
in  the  interior  of  D. 


7.4  I  Exercises  521 


9.  (a)  If  a,  p,  and  k  are  constant  and  the  temperature  T 
of  the  solid  D  is  independent  of  time,  show  that 
the  (net)  heat  flux  of  H  across  the  boundary  of  D 
must  be  zero. 

(b)  Let  D  be  the  solid  region  between  two  concentric 
spheres  of  radii  1  and  2.  Suppose  that  the  inner 
sphere  is  heated  to  120°  C  and  the  outer  sphere  to 
20°  C.  Use  the  result  of  part  (a)  to  describe  the  rate 
of  heat  flow  across  the  spheres. 


1 0.  Consider  the  three-dimensional  heat  equation 


V2u 


du 

a7 


(30) 


for  functions  u(x,  y,  z,  t).  (Here  V2m  denotes  the 
Laplacian32«/3x2  +  d2u/dy2  +  d2u/dz2.)  In  this  ex- 
ercise, show  that  any  solution  T(x,  y,  z,  t)  to  the  heat 
equation  is  unique  in  the  following  sense:  Let  D  be  a 
bounded  solid  region  in  R3  and  suppose  that  the  func- 
tions a(x,  y,  z)  and  <p(x,  y,  z,  t)  are  given.  Then  there 
exists  a  unique  solution  T(x,  y,  z,  t)  to  equation  (30) 
that  satisfies  the  conditions 


T(x,  y,  z,  0)  =  a(x,  y,  z)    for  (x,  y,  z)  €  D, 


and 


(31) 


T(x,  y,  z,  t)  =  4>(x,  y,  z,  t)     for  (x,  y,  z)  6  3D 

and  t  >  0. 

To  establish  uniqueness,  let  T\  and  T2  be  two  solutions 
to  equation  (30)  satisfying  the  conditions  in  (31)  and 
set  w  =  T\  —  T2. 

(a)  Show  that  w  must  also  satisfy  equation  (30),  plus 
the  conditions  that 

w(x,  y,  z,  0)  =  0    for  all  (x,  y,  z)  6  D, 

and 

w(x,  y,  z,t)  =  0    for  all  (x,  y,  z)  e  3D  and  t  >  0. 

(b)  For  t  >  0,  define  the  "energy  function" 
1 


E{t). 


y,z,t)YdV. 


Use  Green's  first  formula  in  Theorem  4.1  to  show 
that  E'(t)  <  0  (i.e.,  that  E  does  not  increase  with 
time). 

(c)  Show  that  E(t)  =  0  for  all  t  >  0.  (Hint:  Show  that 
£(0)  =  0  and  use  part  (b).) 

(d)  Show  that  w(x,  y,  z,  t)  =  0  for  all  t  >0  and 
(x,  y,  z)  e  D,  and  thereby  conclude  the  unique- 
ness of  solutions  to  equation  (30)  that  satisfy  the 
conditions  in  (31). 

11.  Show  that  Ampere's  law  and  Gauss's  law  imply  the 
continuity  equation  for  J.  (Note:  In  the  text,  we  use  the 
continuity  equation  to  derive  Ampere's  law.) 


1 2.  Suppose  that  E  is  an  electric  field,  in  particular,  a  vector 
field  that  satisfies  the  equation  V  •  E  =  p/eo-  A  region 
D  in  space  is  said  to  be  charge-free  if  p  is  zero  at 
all  points  of  D.  Describe  the  charge-free  regions  of 


{x1 


)i+      j  +  (|z3  -  2z)  k. 


13.  By  considering  the  derivation  of  Gauss's  law  for  elec- 
tric fields,  show  that,  for  any  inverse  square  vector 
field  F(x)  =  fcs/||x||3,  the  flux  of  F  across  a  piecewise 
smooth,  closed,  oriented  surface  S  is 

'  dS  if  5  does  not  enclose  the  origin, 

is  [4jik    if  S  encloses  the  origin. 

1 4.  Let  a  point  charge  of  q  coulombs  be  placed  at  the  origin. 
Recover  the  formula 

a  x 

E  =  —  r 

47T60  ||x|| 

by  using  Gauss's  law  in  the  following  way: 

(a)  First,  explain  that  in  spherical  coordinates,  E(x)  = 
E(x)ep,  that  is,  that  E  has  no  components  in  either 
the  ev-  or  ee-direction.  Next,  note  that  E(x)  may 
be  written  as  E(p) — that  is,  that  [|E[|  has  the  same 
magnitude  at  all  points  on  a  sphere  centered  at  the 
origin. 

(b)  Show,  using  Gauss's  law  and  Gauss's  theorem,  that 


Ms 


E(p)ep-dS= 

s  6o 

where  S  is  any  smooth,  closed  surface  enclosing 
the  origin. 

(c)  Now  let  S  be  the  sphere  of  radius  a  centered  at 
the  origin.  Then  the  outward  unit  normal  n  to  S  at 
(p,  <p,  6)  is  ep.  Show  that 


E(p)dS 


(d)  Use  part  (c)  to  show  that  E(p) 
Conclude  the  result  desired. 


q/(4jte0p2). 


15.  (a)  Establish  the  following  identity  for  vector  fields  F 
of  class  C2: 

V  x  (V  x  F)  =  V(V  •  F)  -  V2F. 

(Note:  V2F  =  (V  •  V)F.) 

(b)  In  free  space  (i.e.,  in  the  absence  of  all  charges  and 
currents),  use  Maxwell's  equations  to  show  that  E 
and  B  satisfy  the  wave  equation 
32, 


V2F 


9ZF 


where  k  is  a  constant.  What  is  k  in  each  case? 

(c)  Use  part  (a),  Faraday's  law,  and  Ampere's  law  to 
show  that 


V(V-E)-(V-V)E 


-Mo 


J  +  eo 


3E 

97 


Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


(d)  Show  that,  in  the  absence  of  any  charges  (i.e.,  if 
P  =  0), 


V2E 


9J 


32E 


M0  97  +M06°3^- 


16.  Verify  that  if  the  nonstatic  version  of  Ampere's  law 
(equation  (24))  holds,  then  V  •  (V  x  B)  =  0. 

17.  When  Maxwell  postulated  the  existence  of  displace- 
ment currents  to  arrive  at  a  nonstatic  version  of 
Ampere's  law,  he  was  simply  choosing  the  simplest 
way  to  correct  equation  (22)  so  that  it  would  be  consis- 
tent with  the  continuity  equation  (23).  However,  other 
possibilities  are  also  consistent  with  the  continuity 
equation. 

(a)  Show  that  in  order  to  have  equation  (22)  valid  in 
the  static  case,  then,  in  general,  we  must  have 

3Fi 


V  x  B  =  hqJ  + 


dt 


for  some  (time-varying)  vector  field  Fi  of  class  C2 . 

(b)  By  taking  the  divergence  of  both  sides  of  the  equa- 
tion in  part  (a),  show  that 

3F]  9E 

V  •         =  uofoV  •  —  • 

dt  dt 

(c)  Use  part  (b)  to  argue  that,  from  an  entirely  math- 
ematical perspective,  Ampere's  law  can  also  be 
generalized  as 

9E 

V  x  B  =  moJ  +  Mofo—  +  F2, 
at 

where  F2  is  any  divergence-free  vector  field.  Since 
no  one  has  observed  any  physical  evidence  for  F2 's 
being  nonzero,  it  is  assumed  to  be  zero,  as  in  equa- 
tion (24). 

18.  Suppose  that  J  =  crE.  (This  is  a  version  of  Ohm's  law 
that  obtains  in  some  electric  conductors — here  a  is  a 
positive  constant  known  as  the  conductivity. )  If  p  =  0, 
show  that  E  and  B  satisfy  the  so-called  telegrapher's 
equation, 


9F 


a2F 


V  F  =  moo"  —  +  At06o-  , 
of  at 


1 9.  Let  E  and  B  be  steady-state  electric  and  magnetic  fields 
(i.e.,  E  and  B  are  constant  in  time).  The  Poynting  vec- 
tor field  P  =  E  x  B  represents  radiation  flux  density. 
Use  Maxwell's  equations  to  show  that,  for  a  smooth, 
orientable,  closed  surface  S  bounding  a  solid  region  D, 


E-JdV. 


20.  Consider  the  electric  field  E(r)  defined  by  equation 
(10).  Note  that  the  integrals  in  equation  (10)  are  im- 
proper in  the  sense  that  they  become  infinite  at  points 
r  e  D,  where  p(r)  is  nonzero.  In  this  exercise,  you  will 
show  that,  nonetheless,  the  integrals  in  equation  (10) 
converge  when  D  is  a  bounded  region  in  R3  and  p  is  a 
continuous  charge  density  function  on  D. 

(a)  Write  E(r)  in  terms  of  triple  integrals  for  the  in- 
dividual components.  Let  r  =  (r\,  r2,  r3)  and  x  = 

{x,  y,  z). 

(b)  Show  that  if  each  component  of  E  is  written  in  the 


form  fffpf(x)dV,  then  \f(x) 
where  K  is  a  positive  constant. 

(c)  It  follows  from  part  (b)  that  if 


<  Ki 


IIL 


K 


dV 


converges,  so  must  fffD  f(x)dV,  Show  that 


///„ 


K 


dV 


converges  by  considering  an  iterated  integral  in 
spherical  coordinates  with  origin  at  r.  (Hint:  Look 
carefully  at  the  integrand  in  spherical  coordinates.) 

21.  Consider  the  magnetic  field  B(r)  defined  by  equation 
(15).  As  was  the  case  with  the  electric  field  in  equation 
(10),  it  is  not  obvious  that  the  integrals  in  (15)  converge 
at  all  r  e  D.  Follow  the  ideas  of  Exercise  20  to  show 
that  B(r)  is,  in  fact,  well-defined  at  all  r,  assuming  a 
continuous  current  density  field  J  and  bounded  region 
DinR3. 


True/False  Exercises  for  Chapter  7 


1.  The  function  X:  R2  -*  R3  given  by  X(.s,  t)  =  (2s  + 
3r  +  1 ,  4.?  —  t ,  s  +  2t  —  7)  parametrizes  the  plane 
9x  -  y  -  Uz  =  107. 

2.  The  function  X:  R2  -+  R3  given  by  X(.s,  t )  =  (s2  + 
3t  —  1,  .s2  +  3,  —2s2  +  t)  parametrizes  the  plane  x  — 
ly  -  3z  +  22  =  0. 


3.  The  function  X:(-oo,  oo)  x  (-f ,  f)  ->  R3  given 
by  X(.s,  t)  =  (s3  +3tanr  -  1,  j3  +3,  -2s3  +tanr) 
parametrizes  the  plane  x  —  ly  —  3z  +  22  =  0. 

4.  The  surface  X(s,  t)  =  (s2t,  st2,  st)  is  smooth. 


Miscellaneous  Exercises  for  Chapter  7 


5.  The  area  of  the  portion  of  the  surface  z  =  xexy  lying 
over  the  disk  of  radius  2  centered  at  the  origin  is  given 
by 

\  I  y/l  +  e2xy(x4  +  x2y2  +  2xy  +  \)dy  dx. 
Jo  Jo 

6.  If  S  is  the  unit  sphere  centered  at  the  origin,  then 

ffsx3dS  =  0. 

7.  If  S  is  the  cube  with  the  eight  vertices  (±1,  ±1,  ±1), 
then  ffs(l  +  x3y)dS  =  0. 

8.  If  S  denotes  the  rectangular  box  with  faces  given 
by  the  planes  x  =  ±1,  y  =  ±2,  z  =  ±3,  then 


ffsxyzdS 


0. 


9.  If  S  denotes  the  sphere  of  radius  a  centered  at  the 
origin,  then 

jj(z3-z  +  2)dS  =  j j(x-y5  +2)dS. 

10.  JJS(— yi  +  xj)  •  dS  =  0,  where  S  is  the  cylinder 

9,0  <  z  <  5. 


X2  +  V2 


11.  Let  S  denote  the  closed  cylinder  with  lateral  sur- 
face given  by  y2  +  z2  =  4,  front  by  x  =  7,  and  back 
by  x  =  —  1,  and  oriented  by  outward  normals.  Then 
fsxi-  dS  =  24jt. 

12.  If  S  is  the  portion  of  the  cylinder  x2  +  y2  =  16, 
-2  <  z  <  7,  then  //sVx  (yi)  -dS  =  0. 

13.  ffsF  -  dS  =  6jt,  where  S  is  the  closed  hemisphere 
x2  +  y2  +  z2  =  1,  z  >  0,  together  with  the  surface 
jc2  +  y2  <  1,  z  =  0  and  F  =  yz  i  ■ 


■  *z  i  +  3  k. 


1 9.  If  F  is  a  constant  vector  field,  then  §s  F  •  d S  =  0,  where 
5  is  any  piecewise  smooth,  closed,  orientable  surface. 

20.  §s  V  x  F  •  <iS  =  0,  where  S  is  any  closed,  orientable, 
smooth  surface  in  R3  and  F  is  of  class  C1 . 

21.  Suppose  that  F  is  a  vector  field  of  class  C1  whose  do- 
main contains  the  solid  region  D  in  R3  and  is  such  that 
||F(.v,  y,  z)\\  <  2  at  all  points  on  the  boundary  surface 


S  of  D.  Then 
of  S. 


V  •  F  dV  is  twice  the  surface  area 


22. 


23. 


24. 


25. 


26. 


27. 


28. 


29. 


14.  If  S  is  the  level  set  of  a  function  /(x ,  y ,  z)  and  V  /  ^  0, 
then  the  flux  of  V/  across  S  is  never  zero. 

15.  A  smooth  surface  has  at  most  two  orientations. 

16.  A  smooth,  connected  surface  is  always  orientable. 

17.  If  F  is  a  vector  field  of  class  C1  and  S  is  the  ellipsoid 
x2  +  Ay2  +  9z2  =  36,  then  |/sVxF-dS  =  0. 

18.  1/jV  xF-rfS  has  the  same  value  for  all  piecewise 
smooth,  oriented  surfaces  S  that  have  the  same  bound- 
ary curve  C. 


Miscellaneous  Exercises  for  Chapter  7 


If  S  is  an  orientable,  piecewise  smooth  surface  and  F 
is  a  vector  field  of  class  C1  that  is  everywhere  tangent 
to  the  boundary  of  S,  then  ffs  V  x  F  •  dS  =  0. 

If  S  is  an  orientable,  piecewise  smooth  surface  and  F 
is  a  vector  field  of  class  C1  that  is  everywhere  perpen- 
dicular to  the  boundary  of  S,  then  ffs  V  x  F  •  dS  =  0. 

If  F  is  tangent  to  a  closed  surface  S  that  bounds  a  solid 
region  D  in  R3,  then  jjjDV  -FdV  =  0. 

Let  S  be  a  piecewise  smooth,  orientable  surface  and 
F  a  vector  field  of  class  C1.  Then  the  flux  of  F  across 
S  is  equal  to  the  circulation  of  F  around  the  boundary 
of  S. 

Let  D  be  a  solid  region  in  R3  and  F  a  vector  field  of 
class  C1 .  Then  the  flux  of  F  across  the  boundary  of  D 
is  equal  to  the  integral  of  the  divergence  of  F  over  D. 

Suppose  that  /  and  g  are  of  class  C2  and  D  is  a  solid 
region  in  R3  with  piecewise  smooth  boundary  surface 
S  that  is  oriented  away  from  D.  If  g  is  harmonic,  then 
fffDVf-VgdV  =  §sfVg-dS. 

Suppose  that  /  and  g  are  of  class  C2  and  D  is  a  solid 
region  in  R3  with  piecewise  smooth  boundary  sur- 
face S  that  is  oriented  away  from  D.  If  /  and  g  are 
harmonic,  then  §s  fVg  •  dS  =  —  §s  gV/  •  dS. 

If  V2/  is  known,  then  /  is  uniquely  determined  up  to 
a  constant. 


30.  If  S  is  a  closed,  orientable  surface,  then 


■dS  =  0. 


Figure  7.61  shows  the  plots  of  six  parametrized  sur-  /  / 

faces  X.  Match  each  parametric  description  with  the  X(j'  f)  =  \cost  +  0At  \cost  coss  + 

correct  graph.  j               \  / 

(a)  X(s,t)  =  (t(s2-t2),s,s2-t2)  ^  sin?  sin.?J,sinr  +  0.1r^sinr  coss 

(b)  X(s,  t)  =  (scost,  ssinf,  s)  1              \    t  0.2t 

 -=  cos  t  sin  s  )  ,  — I  —  sin  s 

(c)  X(j,  t)  =  ((2  +  cos  j)cos/,  (2  +  coss)sinf,  r  +  sins)  V5             /    2  V5 

(d)  X(s,    =  (sin*  cos?,  s,  sins  sinf)  (f)  \(s,  t)  =  (s  cost,  s  sin?,  t) 


524       Chapter  7  |  Surface  Integrals  and  Vector  Analysis 


D 


Figure  7.61  Figures  for  Exercise  1. 


2.  Consider  the  unit  sphere  S  with  equation  x2  +  y2  + 
z2  =  1 .  In  this  problem,  you  will  provide  a  parametriza- 
tion  for  (almost  all  of)  the  sphere  that  is  different  from 
the  one  given  in  Example  2  of  §7.1. 
(a)  First  consider  the  parametrized  plane  in  R3 : 

D  =  {(s,  t,  0)  ]  (s,  t)  G  R2}. 

Note  that  D  is  just  a  copy  of  R2  sitting  in  R3 
as  the  jcv-plane.  For  any  point  (s,  t,  0)  G  D,  ar- 
gue geometrically  that  the  line  through  (s,  t ,  0)  and 
(0,  0,1)  intersects  S  in  a  point  other  than  (0,  0,  1). 
(See  Figure  7.62.) 


(b)  Now  define  X:  R2  -+  R3  by  letting  X(s,  t)  be  the 
point  of  intersection  of  S  and  the  line  joining 
{s,  t,  0)  and  (0,  0,  1).  Write  a  set  of  parametric 
equations  for  the  line  joining  (s,  t,  0)  and  (0,  0,  1) 
and  use  it  to  give  a  formula  for  X(s,  t). 

(c)  Show  that  X(s,  t)  is  a  smooth  parametrization  of 
almost  all  of  S.  What  points  of  S  are  not  included 
in  the  image  of  X?  For  evident  geometric  reasons, 
the  map  X  defined  in  this  problem  has  an  inverse 
map  from  (almost  all  of)  the  sphere  to  the  x  v-plane, 
called  stereographic  projection  of  the  sphere  onto 
the  plane. 

3.  (a)  Provide  a  parametrization  for  the  hyperboloidx2  + 
y2  —  z2  =  1.  (Hint:  Use  the  cylindrical  coordinates 
z  and  0  for  parameters.) 

(b)  Modify  your  answer  in  part  (a)  to  give  a  para- 
metrization of  the  hyperboloid 


(c)  Let  jco  and  yo  be  such  that  x2  +  y2  =  I.  Show  that 
the  lines 


Figure  7.62  Figure  for  Exercise  2. 


ll(0  =  (a(x0  -  y0t),  b(x0t  +  y0),  ct) 


Miscellaneous  Exercises  for  Chapter  7 


and 


h(t)  =  (a(y0t  +  x0),  b(y0  -  x0t),  ct) 


lie  in  the  hyperboloid  of  part  (b). 

(d)  Show  that  the  lines  li  and  I2  of  part  (c)  also  lie  in 
the  plane  tangent  to  the  hyperboloid  at  the  point 
(ax0,  by0,  0). 


4.  Find  the  surface  area  of  the  portion  of  the  hyperboloid 
x2  +  y2  —  z2  =  1  between  z  =  —  a  and  z  =  a.  (See 
Exercise  3(a).) 


5.  (a)  Parametrize  the  ellipsoid 


xl  z!  zl 

a2  +  b2+  c2 


1. 


(b)  Use  your  answer  in  part  (a)  to  set  up  an  integral 
for  the  surface  area  of  the  ellipsoid.  Do  not  evaluate 
this  integral,  but  verify  that  it  indicates  correctly 
the  surface  area  in  the  case  that  the  ellipsoid  is  a 
sphere  of  radius  a. 

6.  Let  y  =  f(x),  a  <  x  <  b,  be  a  curve  in  the  xy-plane. 
Suppose  this  curve  is  revolved  around  the  x-axis  to 
generate  a  surface  of  revolution. 

(a)  Explain  why  X(s,  t)  =  (s,  f(s)  cos  t,  f(s)  sint) 
parametrizes  the  surface  so  described.  (Hint:  Con- 
sider the  r-coordinate  curve.) 

(b)  Verify  that  the  area  of  the  surface  is 


2n 


/  \f(X)\y/l  +  {f'(x))2dx. 
J  a 


7.  Let  the  curve  y  =  f(x),  0  <  a  <  x  <  b,  be  revolved 
around  the  v-axis  to  generate  a  surface. 

(a)  Find  a  parametrization  for  the  surface. 

(b)  Verify  that  the  area  of  the  surface  is 


In  \   xjl+(f'(x))2  dx. 

J  a 


8.  Let  S  denote  the  surface  defined  by  the  equation 
z  =  f(x),  a  <  x  <  b.  (The  surface  is  a  generalized 
cylinder  over  the  curve  z  =  f(x)  in  the  xz-plane.)  Let 
C  denote  a  piecewise  C1,  simple,  closed  curve  in  the 
.icy-plane.  Let  D  denote  the  region  in  the  xy-plane 
bounded  by  C  and  assume  that  every  point  of  D  has  x- 
coordinate  between  a  and  b.  Let  Si  denote  the  portion 
of  S  lying  over  D. 

(a)  Show  that  the  portion  of  S  lying  over  D  is 


s'(x)dA, 


where  s(x)  =  f*  ^1  +  (f'(t))2  dt ;  that  is,  s  is  the 
arclength  function  of  the  curve  z  =  f(x). 


(b)  Use  part  (a)  to  show  that  the  surface  area  may  also 
be  calculated  from  the  line  integral 


I 


s(x)dy, 


where  C  is  oriented  counterclockwise, 
(c)  Compute  the  surface  area  of  the  portion  of 


x>  1 
Z  =  —  +  — . 
3  4x 


1<  x  <  3, 


lying  over  the  rectangle  in  the  xy-plane  having  ver- 
tices (1,  ±2),  (3,  ±2). 

Let  f  be  a  piecewise  smooth  surface  in  R3  and  f:X  C  R3  — >• 
R  a  scalar-valued  function  whose  domain  X  contains  S.  Then 
the  average  value  of  f  on  S  is  the  quantity 

[fU  =  Sfsfds  =  SfsfdS 

ffsdS       area  of  S 

Exercises  9—11  involve  the  notion  of  the  average  value  of  a 
function  on  a  surface. 

9.  (a)  Explain  why  the  definition  of  the  average  value 
makes  sense. 

(b)  Suppose  that  the  temperature  at  points  on  the 
sphere  x2  +  y2  +  z2  =  49  is  given  by  T(x,  y,  z)  = 
x2  +  y2  —  3z.  Find  the  average  temperature. 

10.  Find  the  average  value  of  f(x,  y,  z)  =  x2ez  —  y2z  on 
the  cylinder  x2  +  y2  =  4,  0  <  z  <  3. 

1 1 .  Find  the  average  value  of  f(x ,  y ,  z)  =  x2  +  y2  —  3  on 
the  portion  of  the  cone  z2  =  4x2  +  Ay2,  —2  <  z  <  6. 

1 2.  A  thin  film  is  made  in  the  shape  of  the  helicoid 

X(s,  t)  =  (s  cost,  ssint,t),     0  <  s  <  1,  0  <  t  <  4ir. 

Suppose  that  the  mass  density  (per  unit  area)  at  each 
point  (x,  y,  z)  of  the  film  varies  as 

S(x,  y,  z)  =  Jx2  +  y2. 

Find  the  total  mass  of  the  film. 

Let  S  be  a  piecewise  smooth  surface  in  R3.  Suppose  that  the 
mass  density  at  points  (x,  y,  z)  of  S  is  S(x ,  y ,  z).  Using  formu- 
las analogous  to  those  in  §5.6,  we  define  the  (first)  moments 
of  S  to  be 


x8(x,  y,  z)dS, 


SL 


yS(x,  y,  z)dS, 


and 


j j  zS(x,  y,  z)dS. 


Exercises  13—16  involve  first  moments  and  centers  of  mass  of 
surfaces. 

1 3.  Find  the  center  of  mass  of  the  first  octant  portion  of  the 
sphere 

x2  +  y2  +  z2  =  a2, 
assuming  constant  density. 


Chapter  7  I  Surface  Integrals  and  Vector  Analysis 


14.  Find  the  center  of  mass  of  the  piece  of  the  cylinder 

x2  +  z2=a2,  0<y<a,z>0. 
Assume  that  the  density  of  the  surface  is  constant. 

15.  Find  the  center  of  mass  of  a  sphere  of  radius  a,  where 
the  density  S  varies  as  the  square  of  the  distance  from 
the  "south  pole"  of  the  sphere. 

16.  Find  the  center  of  mass  of  the  cylinder  x1  +  z2  =  a2, 
between  y  =  0  and  y  =  2,  if  the  density  varies  as 

8  =  x2  +  y. 

Given  a  piecewise  smooth  surface  S,  the  moment  of  inertia  I- 
of  S  about  the  z-axis  is  defined  by  the  surface  integral 


I- 


j j  (x2  +  y2)SdS, 


where  8(x,  y,  z)  is  mass  density.  The  corresponding  radius  of 
gyration  about  the  z-axis  is  given  by 


where  M 


,SdS.  Likewise,  the  moments  of  inertia  of  S 


about  the  x-  and  y-axes  are  given  by,  respectively,  the  surface 


fjy+z2)sds 


and 


I,  =  j j(x2  +  z2)SdS. 


(Compare  these  formulas  with  the  ones  in  §5.6.)  Exercises  17— 
19  concern  moments  of  inertia  andradii  of  gyration  of  surfaces. 

1 7.  (a)  Calculate  /-  for  the  surface  S  cut  from  the  cone 

z2  =  4x2  +  Ay2 

by  the  planes  z  =  2  and  z  =  4.  (This  surface  is 
known  as  a  frustum  of  the  cone.)  Assume  density 
is  equal  to  1 . 

(b)  Find  the  radius  of  gyration  r.  of  the  frustum. 

(c)  Repeat  parts  (a)  and  (b),  assuming  that  the  density 
at  a  point  is  proportional  to  the  distance  from  that 
point  to  the  axis  of  the  cone. 

18.  Let  S  denote  the  cylindrical  surface  with  equation 
x2  +  y2  =  a2,  where  —  b  <  z  <  b  (a,  b  positive  con- 
stants). Assume  that  the  density  S  is  constant  along 
S. 

(a)  Find  the  moment  of  inertia  of  S  about  the  z-axis. 

(b)  Find  the  radius  of  gyration  of  S  about  the  z-axis. 

19.  Let  S  be  as  in  Exercise  18. 

(a)  Find  the  moments  of  inertia  Ix  and  Iy  of  S  about 
the  x-  and  y-axes. 

(b)  Find  the  radii  of  gyration  rx  and  ry . 

20.  (a)  Prove  the  following  mean  value  theorem  for  double 

integrals:  If  /  and  g  are  continuous  on  a  compact, 


connected  region  D  C  R2 ,  then  there  is  some  point 
P  e  D  such  that 

J f  fgdA  =  f(P)  jj^dA. 

(Hint:  Consider  the  ratio  JfDfgdA/  JJDgdA 
and  use  the  intermediate  value  theorem.) 

(b)  Use  the  result  of  part  (a)  to  prove  the  following: 
Let  S  be  a  smooth  surface  oriented  by  unit  normal 
n  and  let  F  be  a  continuous  vector  field  on  S.  As- 
sume that  S  may  be  parametrized  by  a  single  map 
X:  D  — >  R3 .  Then  there  is  some  point  P  e  S  such 
that 


F-dS=  [F(P)-n(P)](area  of  S). 


21.  Let  a  be  a  constant  vector  and  C  a  smooth,  simple, 
closed  curve.  Show  that  j>c  a  •  ds  =  0  in  two  ways: 

(a)  directly; 

(b)  by  assuming  that  C  is  the  boundary  of  a  smooth 
surface  S. 

22.  Evaluate  §c(x2  +  z2)dx  +  y  dy  +  z  dz,  where  C  is 
the  closed  curve  parametrized  by  the  path  x(f)  = 
(cos  t ,  sin  t ,  cos2  t  —  sin2  r). 

23.  Let  /  and  g  be  functions  of  class  C2,  and  let  S  be  a 
piecewise  smooth,  orientable  surface.  Show  that 

(JVg)-d»=  f  f(VfxVg).dS. 

as  J  Js 

24.  Let  /  and  g  be  functions  of  class  C2,  and  let  S  be  a 
piecewise  smooth,  orientable  surface.  Show  that 


(fVg  +  gVf)-ds  =  0. 


(Hint:  Use  Exercise  23.) 

25.  Let  /  be  of  class  C2,  and  let  S  be  a  piecewise  smooth, 
orientable  surface.  Show  that 


Jdi 


(/V/)-</s  =  0. 


26.  Let  C  be  a  simple,  closed,  piecewise  C1  planar  curve 
in  R3 .  That  is,  C  is  contained  in  some  plane  in  R3 .  Let 
n  =  ai  +  foj  +  ck  denote  a  unit  vector  normal  to  the 
plane  containing  C,  and  let  C  be  oriented  by  a  right- 
hand  rule  with  respect  to  n. 

(a)  Show  that 

1  r 

-  (t  (bz  —  cy)dx  +  (cx  —  az)dy  +  (ay  —  bx)dz 

2  Jc 

=  area  enclosed  by  C. 

(b)  Now  show  that  the  result  of  part  (a)  reduces  to 
something  familiar  in  the  case  that  C  is  a  curve  in 
the  Jcy-plane.  (See  Example  2  of  §6.2.) 


Miscellaneous  Exercises  for  Chapter  7 


27.  Suppose  S  is  a  piecewise  smooth,  orientable  surface 
with  boundary  dS.  Use  Faraday's  law  to  show  that  if  E 
is  everywhere  perpendicular  to  dS,  then  the  magnetic 
flux  induced  by  E  does  not  vary  with  time. 

28.  Let  G  be  a  vector  field  of  class  C2.  Then  by  Theorem 
4.4  in  Chapter  3,  V  •  (V  x  G)  =  0.  Therefore,  Gauss's 
theorem  implies  that 

jj(VxG)-dS  =  jjj  V-(V  xG)dV  =  0. 

Then  Stokes's  theorem  yields 

0=  f  f(VxG)^S=l  G-ds. 

J  J  S  J  95 

Hence,  all  vector  fields  G  of  class  C2  are  conservative. 
How  can  this  be? 

Recall  that  we  measure  an  angle  in  radians  as  follows:  Place 
the  vertex  of  the  angle  at  the  center  O  of  a  circle  of  radius  a, 
so  that  the  angle  subtends  an  arc  of  the  circle  of  length  s.  Then 
the  measure  of  the  angle  in  radians  is 


s 


We  can  do  something  similar  in  the  three-dimensional  case  by 
defining  a  solid  angle  as  the  set  of  rays  beginning  at  the  center 
O  of  a  sphere  of  radius  a,  so  that  the  rays  cut  out  a  portion 
of  the  sphere  having  surface  area  A.  Then  the  measure  of  the 
solid  angle  in  steradians  is 


Figure  7.63  A  solid  angle 
measured  on  a  sphere. 


Now  suppose  that  S  is  a  smooth,  oriented  surface  and  that  a 
point  O  in  R3  is  chosen  that  is  not  in  S  and  so  that  every  line 
through  O  intersects  S  at  most  once.  In  this  case,  we  define 
the  solid  angle  relative  to  O  subtended  by  S  as  the  set  of  rays 
beginning  at  O  that  pass  through  a  point  of  S.  We  measure  this 


solid  angle  by  calculating 

£i(S,0)  =  JJ^^.dS,  (1) 

where  x  denotes  the  (varying)  vector  from  O  to  a  point  P  in  S 
and  S  is  oriented  by  a  normal  that  points  away  from  O. 

Exercises  29—32  develop  some  ideas  regarding  solid  angles. 

29.  In  this  problem  we'll  see  how  the  measure  of  the 
solid  angle  given  in  equation  (1)  can  be  related  to  the 
more  geometric  notion  of  measuring  solid  angles  using 
spheres.  To  that  end,  in  the  situation  described  above, 
construct  a  sphere  Sa  of  radius  a  centered  at  O .  Let  Sa 
denote  the  intersection  of  S„  and  the  solid  angle  of  S 
relative  to  O.  (See  Figure  7.64.)  By  applying  Gauss's 
theorem  to  the  solid  region  W  between  S  and  Sa ,  show 
that 


surface  area  of  St 


Figure  7.64  A  solid  angle  subtended  by  a 
surface  S. 


30.  If  S  is  parametrized  by  X:  D  C  R2  -»•  R3,  and  O  de- 
notes the  origin  in  R3 ,  we  may  use  equation  ( 1 )  to  define 
the  measure  of  the  solid  angle  subtended  by  S  for  very 
general  surfaces,  including  ones  that  may  have  some 
self-intersections.  Show  that  for  such  a  parametrized 
surface        O)  may  be  calculated  as 


-  X 

y 

z  ~ 

dx 

dy 

dz 

~ds~ 

ds 

~dl 

dx 

dy 

dz 

-  ~dt 

dt 

dt  - 

31 .  Suppose  that  S  is  closed,  oriented,  and  forms  the  bound- 
ary of  a  solid,  bounded,  simply-connected  region  W  in 


Chapter  7  I  Surface  Integrals  and  Vector  Analysis 


R3.  Using  equation  (1)  to  define  the  measure  of  the 
solid  angle,  show  that 

f  ±4jr    if  S  encloses  O 


38.  Suppose  that 


Q{S,  O) 


lo 


if  S  does  not  enclose  O. 


32.  Suppose  that  S  is  the  circular  disk  of  radius  a  in  the  xy- 
plane  and  centered  at  the  origin.  Let  O  be  a  moving 
point  along  the  z-axis  and  denote  it  by  (0,  0,  z).  Use 
equation  (1)  to  show  that  — 2jt  <  f2(S,  O)  <  2n.  In 
addition,  show  that  Q(S,  O)  jumps  by  An  as  O  passes 
through  S. 

33.  Prove  the  following:  Let  F  be  a  vector  field  of  class 
C1  defined  on  R3.  If  V  •  F  =  0,  then  there  is  some 
vector  field  G  of  class  C1  such  that  F  =  V  x  G.  This 
result  provides  a  converse  to  Theorem  4.4  of  Chapter  3 . 
(Hint:  Let 

G(x,y,z)=  /    t¥(xt,  yt,  zt)  x  rdt, 
Jo 

where  r  =  x  i  +  y  j  +  z  k.  Show  V  x  G  =  F.  The 
identities 

V  x  (A  x  B)  =  AV  •  B  -  BV  •  A 

+  (B-V)A-(A- V)B, 


dt 


[t¥(xt,  yt,  zt)]  =  (r  •  VjF(xt,  yt,  zt) 
+  F(xt,  yt,  zt), 


and 


d 

dt 


[t2F(xt,  yt,  zt)]  =  t  |-^|)F(jef,  yt,  zt)] 

'I 


+  F(xt,yt,zt)\ 


will  prove  useful.  Also,  let  X  =  xt,  Y  =  yt,  Z  =  zt 
and  note  that,  by  the  chain  rule, 


3F  _  9F  dX  _  9F 

~dx  ~  dxJx  ~  'dX' 


etc.) 


The  vector  field  G  defined  in  Exercise  33  is  called  a  vector 
potential  for  the  vector  field  F.  In  Exercises  34—36,  determine 
a  vector  potential  for  the  given  vector  field  or  explain  why  such 
a  potential  fails  to  exist. 

34.  F  =  2xi-yj-zk 

35.  F  =  xi  +  yj+zk 

36.  F  =  2>y\  +  2xz\ -  7x2yk 

37.  The  vector  potential  G  identified  in  Exercise  33  is  not 
unique.  In  fact,  show  that  if  G  is  a  vector  potential  for  F, 
then  so  is  G  +  V0,  where  <fr  is  any  scalar-valued  func- 
tion of  class  C2 .  (This  is  known  as  the  gauge  freedom 
in  choosing  the  vector  potential.) 


GMm 


is  the  gravitational  force  field  defined  on  R3  — 
{(0,0,0)}. 

(a)  Show  that  V  •  F  =  0  by  direct  calculation. 

(b)  Show  that  F  /  V  x  G  for  any  C1  vector  field  G 
defined  on  R3  -  {(0,  0,  0)}  by  using  Stokes's  the- 
orem. (Hint:  Take  a  sphere  S  enclosing  the  origin 
and  break  it  up  into  the  upper  and  lower  hemi- 
spheres. Consider  ffs  F  •  dS  as  the  sum  of  the  sur- 
face integrals  over  the  two  hemispheres.) 

(c)  Why  do  parts  (a)  and  (b)  not  contradict  the  result 
of  Exercise  33? 

In  Exercises  39—44  below,  you  will  derive  a  type  of  wave  equa- 
tion and  see  how  Maxwell  s  equations  can  be  reduced  to  this 
wave  equation.  Assume  that  the  electric  and  magnetic  fields  E 
and  B  are  defined  on  a  simply-connected  region  in  R3;  also 
let  the  symbol  V  denote  differentiation  with  respect  to  x,y,z 
(i.e.,  not  with  respect  to  t). 

39.  Recall  that  equation  (20)  in  §7.4  was  derived  by  show- 
ing that  the  magnetic  field  B  had  a  vector  potential 
A  of  class  C1  (see  Exercise  33  above);  that  is,  that 
B  =  V  x  A.  Use  Faraday's  law  (equation  (27))  to  show 
that  the  vector  field  E  +  dA/dt  is  conservative.  Hence, 
we  may  write 

9A 

E  +  ^=V/ 
for  an  appropriate  scalar-valued  function  f(x,  y,  z,  t). 

40.  Use  the  vector  potential  A  for  B  in  Ampere's  law  (equa- 


tion (24)  of  §7.4)  to  conclude  that 
32A 


dt2 


Ho  J  +  V   V  •  A  -  iA,0e0 


(2) 


(Hint:  See  part  (a)  of  Exercise  15  of  §7.4.) 


41.  Use  Gauss's  law  (equation  (12)  of  §7.4)  to  show  that 

(3) 


V2/ 


p  9 
60  dt 


42.  As  noted  in  Exercise  37,  the  vector  potential  A  is  only 
unique  up  to  addition  of  V0,  where  (p(x,  y,  z,  t)  is  a 
scalar- valued  function  of  class  C2 .  That  is,  any  vector 
field  A  =  A  +  V0  also  works  as  a  vector  potential  for 
B.  However,  the  function  /  that  arises  in  Exercises  39- 
41  will  change. 

(a)  Show  that  the  function  /  associated  to  A  is  related 
to  /  as 


/  =  /  + 


dt 


Miscellaneous  Exercises  for  Chapter  7 


(b)  Show  that  the  condition 


V  •  A  =  (U,06o 


9/ 
3r  ' 


where  A  =  A  +  V0,  is  equivalent  to  the  existence 
of  solutions  to  the  (inhomogeneous)  wave  equation 


d2<p 


V  •  A  -  ix0eQ 


(4) 


Given  A  and  /  it  is  possible  to  solve  equation 
(4)  for  <p,  so  we  may  assume  that  the  condition 
V  •  A  =  fio€odf/dt  holds. 


43.  Given  the  condition  V  •  A  =  iiQ^df/dt,  show  that 
equations  (2)  and  (3)  become 

92A 


V2A- 


V2/  -  Mofo 


3r2 

S2/ 
9r2 


-/x0J; 


P_ 


(5) 


(6) 


44.  Conversely,  suppose  that  A  and  /  satisfy  the  condition 
V  •  A  =  fioeodf/dt  and  equations  (5)  and  (6).  Show 
that  then  E  =  -dA/dt  +  V/  and  B  =  V  x  A  must 
satisfy  Maxwell's  equations.  Hence,  solutions  to  (4) 
enable  us  to  define  a  vector  field  A  and  a  scalar  field 
/  from  which  we  may  construct  E  and  B  that  satisfy 
Maxwell's  equations. 


Vector  Analysis  in 
Higher  Dimensions 


8.1  An  Introduction  to 
Differential  Forms 

8.2  Manifolds  and  Integrals  of 
k-forms 

8.3  The  Generalized  Stokes's 
Theorem 

True/False  Exercises  for 
Chapter  8 

Miscellaneous  Exercises  for 
Chapter  8 


Introduction 

In  this  concluding  chapter,  our  goal  is  to  find  a  way  to  unify  and  extend  the 
three  main  theorems  of  vector  analysis  (namely,  the  theorems  of  Green,  Gauss, 
and  Stokes).  To  accomplish  such  a  task,  we  need  to  develop  the  notion  of  a 
differential  form  whose  integral  embraces  and  generalizes  line,  surface,  and 
volume  integrals. 

8.1    An  Introduction  to  Differential  Forms 

Throughout  this  section,  U  will  denote  an  open  set  in  R",  where  R"  has  coor- 
dinates (jci,  xz,  ■  ■  ■ ,  xn),  as  usual.  Any  functions  that  appear  are  assumed  to  be 
appropriately  differentiable. 


Differential  Forms   

We  begin  by  giving  a  new  name  to  an  old  friend.  If  /:  U  c  R"  — >•  R  is  a  scalar- 
valued  function  (of  class  Ck),  we  will  also  refer  to  /  as  a  differential  0-form, 
or  just  a  0-form  for  short.  0-forms  can  be  added  to  one  another  and  multiplied 
together,  as  well  we  know. 

The  next  step  is  to  describe  differential  1 -forms.  Ultimately,  we  will  see  that  a 
differential  1-form  is  a  generalization  of  f{x)  dx — that  is,  of  something  that  can 
be  integrated  with  respect  to  a  single  variable,  such  as  with  a  line  integral.  More 
precisely,  in  R",  the  basic  differential  1-forms  are  denoted  dx\,  dxz,  ■  ■ . ,  dx„. 
A  general  (differential)  1-form  a>  is  an  expression  that  is  built  from  the  basic 
1-forms  as 

to  =  Fi(xu       xn)dx\  +  F2(xi,  . . .  ,xn)dx2  +  ■■■  +  Fn(xu  xn)dx„, 

where,  for  j  =  1, . . . ,  n,  Fj  is  a  scalar- valued  function  (of  class  Ck)  on  U  c  R". 
Differential  1-forms  can  be  added  to  one  another,  and  we  can  multiply  a  0-form 
/  and  a  1-form  to  (both  defined  on  U  c  R")  in  the  obvious  way:  If 

co  =  F\  dx\  +  F2  dx2  H  +  Fn  dxn , 

then 


fco  =  fF\  dx\  +  fFz  dx2  H  h  f  Fn  dx„. 


8.1  |  An  Introduction  to  Differential  Forms 


EXAMPLE  1    In  R3,  let 

co  =  xyz  dx  +  z2  cos  y  dy  +  zex  dz  and  >?  =  (y  —  z) dx  +  z2  sin y  dy  —  2dz- 
Then 

a>  +  r]  =  (xyz  +  y  —  z)dx  +  z2(cos  y  +  sin  y)  dy  +  (zex  —  T)dz. 
If  f{x,  y,  z)  =  xey  —  z,  then 

fco  =  (xey  —  z)xyz  dx  +  (xey  —  z)z2  cos  y  dy  +  (xey  —  z.)zexdz.  ♦ 

Thus  far,  we  have  described  1 -forms  merely  as  formal  expressions  in  certain 
symbols.  But  1 -forms  can  also  be  thought  of  as  functions.  The  basic  1 -forms 
dx\, ... ,  dx„  take  as  argument  a  vector  a  =  (a\,  a%, . . . ,  an)  in  R";  the  value  of 
dxt  on  a  is 

dxj  (a)  =  a, . 

In  others  words,  dx,  extracts  the  tth  component  of  the  vector  a. 

More  generally,  for  each  xo  e  U,  the  1-form  co  gives  rise  to  a  combination 
cyXo  of  basic  1 -forms 

tt'xo  =  ^i(xo)^i  H  r-  F„(x0)£/x„; 

coXo  acts  on  the  vector  a  e  R"  as 

«Xo(a)  =  Fi(x0)  Jxj(a)  +  F2(x0)^2(a)  H  h  F„(x0)dx„(a). 

EXAMPLE  2   Suppose  o>  is  the  1-form  defined  on  R3  by 

co  =  x2yz  dx  +  y2z  dy  —  3xyz  dz. 
If  x0  =  (1,  —2,  5)  and  a  =  (a\,  a2,  03),  then 

W(i-2,5)(a)  =  -10<ijr(a)  +  20Jv(a)  +  30<iz(a) 
=  -10«i  +20fl2  +  30fl3, 

and,  if  x0  =  (3,  4,  6),  then 

a>(3  4  6)(a)  =  216<ix(a)  +  96c?v(a)  —  216Jz(a) 

=  216fli  +  96a2  —  216a3. 

The  notation  suggests  that  a  1-form  is  a  function  of  the  vector  a  but  that  this 
function  varies  from  point  to  point  as  xo  changes.  Indeed  1 -forms  are  actually 
functions  on  vector  fields.  ♦ 


A  basic  (differential)  2-form  on  R"  is  an  expression  of  the  form 

dxj  A  dxj ,   i,  j  =  1, . . . ,  n. 

It  is  also  a  function  that  requires  two  vector  arguments  a  and  b,  and  we  evaluate 
this  function  as 


dxt  A  dxj(a,  b)  = 


dxi(a)  dxiQa) 
dxj(&)  dxj(b) 


(The  determinant  represents,  up  to  sign,  the  area  of  the  parallelogram  spanned 
by  the  projections  of  a  and  b  in  the  XjXj  -plane.)  It  is  not  difficult  to  see  that,  for 
i,  j  =  1- 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


dxj  A  dxj  =  —dxj  A  dxj 

(1) 

and 

dxj  A  dxi  =  0. 

(2) 

Formula  (1)  can  be  established  by  comparing  dxj  A  dxj(a,  b)  with  dxj  A 
dx,(a,  b).  Formula  (2)  follows  from  formula  (1).  Given  formulas  (1)  and  (2), 
we  see  that  we  can  generate  all  the  linearly  independent,  nontrivial  basic  2-forms 
on  R"  by  listing  all  possible  terms  dx{  A  dxj,  where  i  and  j  are  integers  between 
1  and  n  with  i  <  j: 

dx\  A  dx2,  dx\  A  dxi, . . . ,  dx\  A  dxn, 
dx2  A  dx-i,  . . . ,  dx2  A  dxn, 


dx„-i  A  dxn. 

To  count  how  many  2-forms  are  in  this  list,  note  that  there  are  n  choices  for  dx{ 
and  n  —  1  choices  for  dxj  (so  that  dxt  ^  dxj  in  view  of  (2)),  and  a  "correction" 
factor  of2so  as  not  to  count  both  dxj  Adxj  anddxj  A  dxj  in  light  of  (1).  Hence, 
there  are  n(n  —  l)/2  independent  2-forms. 

Let  x  =  (xi,  jt2, . . . ,  x„).  A  general  (differential)  2-form  on  U  c  R"  is  an 
expression 

o)  =  F\2{~iL)dx\  A  dx2  +  Fii(x)dxi  A  dx-i  +  •  ■  •  +  Fn_\n(\)dxn^\  A  dxn, 

where  each  Fy  is  a  real-valued  function  Fy :  C/  c  R"  ->  R.  The  idea  here  is  to 
generalize  something  that  can  be  integrated  with  respect  to  two  variables — such 
as  with  a  surface  integral. 

EXAMPLE  3   In  R3,  a  general  2-form  may  be  written  as 

F\(x,  y,  z)dy  Adz  +  Fi{x,  y,  z)dz  A  dx  +  F^(x,  y,  z)dx  A  dy. 

The  reason  for  using  this  somewhat  curious  ordering  of  the  terms  in  the  sum  will, 
we  hope,  become  clear  later  in  the  chapter.  ♦ 

Given  a  point  xo  6  U  c  R",  to  evaluate  a  general  2-form  on  the  ordered  pair 
(a,  b)  of  vectors,  we  have 

wXo(a,  b)  =  Fi2(x0)<ixi  A  dx2(a,  b)  +  Fi3(x0)<ixi  A  Jx3(a,  b) 
+  h  F„_i„(x0)(ix„_i  A  dx„(a,  b). 

EXAMPLE  4  In  R3,  let  co  =  3xy  dy  A  dz  +  (2y  +  z)  dz  A  dx  +  (x  -  z)  dx  A 
dy.  Then 

W(i.2,-3)(a,  b)  =  6  dy  A  dz(&,  b)  +  dz  A  Jx(a,  b)  +  4  dx  A  <iy(a,  b) 


«3 


fc3 


+ 


#3 


&3 
/'I 


+  4 


«2 


&1 
fe2 


=  6(a2^3  -  03^2)  +  (^3^1  -  aibi)  +  4(aiZ?2  -  «2^i)- 


8.1  I  An  Introduction  to  Differential  Forms  533 


Finally,  we  generalize  the  notions  of  1 -forms  and  2-forms  to  provide  a  defi- 
nition of  a  ft-form. 


DEFINITION  1 .1    Let  k  be  a  positive  integer.  A  basic  (differential)  A-form 

on  R"  is  an  expression  of  the  form 

dxj{  A  dxj2  A  ■  ■  ■  A  dxik , 

where  1  <  ij  <  n  for  j  =  1 ,  . . . ,  k.  The  basic  &-forms  are  also  functions  that 
require  k  vector  arguments  ai,  aj, . . . ,  a*  and  are  evaluated  as 

dx^i&i)  dxh{2Lz)  ■■■  dxh{ak) 


dxix  A  ■  ■  ■  A  dxik(n\ , . . . ,  a^)  =  det 


dxi2{n\)  dxi2(a2)  ■  ■  ■  dxj2(ak) 


dxit(a\)  dxit(a2)  ■■■  dxit(ak) 


EXAMPLE  5  Let 

ai  =  (1,2, -1,3,0),  a2  =  (5,  4,  3, 2, 1),  and  a3  =  (0,  1,  3, -2,  0) 
be  three  vectors  in  R5.  Then  we  have 


dx\  A  dxT,  A  dx5(a\,  a2,  a^)  =  det 
Using  properties  of  determinants,  we  can  show  that 


1  5  0 
1  3  3 
0      1  0 


dxi,  A  ■  ■  •  A  dxj.  A  •  •  ■  A  dxi,  A  •  •  ■  A  dXik 

(3) 

=  —dXjl  A  ■  •  •  A  dXjt  A  •  •  •  A  dXjj  A  •  •  •  A  dXik 

and 

dxix  A  ■  •  ■  A  dX[  A  ■  ■  •  A  dXjj  A  ...  A  dXjk  =  0. 

(4) 

Formula  (3)  says  that  switching  two  terms  (namely,  dxi  and  dxt,)  in  the  basic 
&-form  dxix  A  •  •  •  A  dxik  causes  a  sign  change,  and  formula  (4)  says  that  a  basic 
&-form  containing  two  identical  terms  is  zero.  Formulas  (3)  and  (4)  generalize 
formulas  (1)  and  (2). 


DEFINITION  1 .2  A  general  (differential)  A-form  on  U  C  R"  is  an  expres- 
sion of  the  form 

n 

co  =  Fii  ^jk(x)dxii  A  ■  •  ■  A  dxik, 

i'i,...,i't=l 

where  each  f^,,.*,  is  a  real-valued  function  F,,  ..,t:  [7  — >•  R.  Given  a  point 
xo  £  {/,  we  evaluate  to  on  an  ordered  £-tuple  (ai , . . . ,  a*)  of  vectors  as 

n 

coXo(au...,ak)=    ^2   Fh-ikix<y)dxh  A--- Adxik{ax, . . .  ,ak). 

h,...,ik—l 


534       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Note  that  a  0-form  is  so  named  because,  in  order  to  be  consistent  with  a 
1-form  or  2-form,  it  must  take  zero  vector  arguments! 

In  view  of  formulas  (3)  and  (4),  we  write  a  general  /c-form  as 

a>  =       ^2      Fh-k^xh  A  •  •  •  A  dxik. 

\<i\  <---<ifr<n 

(That  is,  the  sum  may  be  taken  over  strictly  increasing  indices  i\, . , . ,  4.)  For 
example,  the  4-form 

co  =  xi  dx\  A  dx?,  A  dx\  A  dx$  +  {xj,  —  x\)dx\  A  dx-±  A  dx$  A  dxi, 
+  X\X^  dx$  A  t/x3  A  dx4  A  Jxi 

may  be  written  in  the  "standard  form"  with  increasing  indices  as 

CO  =  (X2  —  X\Xi)dx\  A  <ix3  A  dx4  A  dxs  +  (xj  —  X3)  dx\  A  dx2  A  dx3  A  dx$. 

Two  &-forms  may  be  added  in  the  obvious  way,  and  the  product  of  a  0-form 
/  and  a  &-form  co  is  analogous  to  the  product  of  a  0-form  and  a  1-form. 

Exterior  Product  

The  symbol  A  that  we  have  been  using  does,  in  fact,  denote  a  type  of  multiplication 
called  the  exterior  (or  wedge)  product.  The  exterior  product  can  be  extended  to 
general  differential  forms  in  the  following  manner: 


DEFINITION  1.3    Lett/  c  R"  be  open.  Let  /  denote  a  0-form  on  U .  Let  co  = 
dx,  A  •  •  •  A  dxik  denote  a  &-form  on  U  and  17  =  J]  Gjl..jldxj1  A 
•  •  •  A  dxj,  an  /-form.  Then  we  define 

/  A  co  =  fco  =       fFh...hdxh  A  ■  •  ■  A  dx,t, 

CO  A  r\  =         Fii...ikG j1...jldXjl  A  •  •  ■  A  <£tjs  A  dXj,  A  ■  ■  •  A  fi?X/;. 

Thus,  the  wedge  product  of  a  &-form  and  an  /-form  is  a  (&  +  /)-form. 

EXAMPLE  6  Let 

co  =  x\  dx\  A  dx2  +  (2x3  —  X2)dx\  A  JX3  +  exi  dxj  A  JX4 

and 

T]  =  X4  Jxj  A  JX3  A  JX5  +  X6  JX2  A  JX4  A  dX(, 

be,  respectively,  a  2-form  and  a  3-form  on  R6.  Then  Definition  1.3  yields 

co  A  rj  =  x\xa,  dx\  A  dx2  A  Jxj  A  JX3  A  dx$ 

+  (2X3  —  X2)X4  dx\  A  JX3  A  dx\  A  C/X3  A  dx$ 

+  eX3X4  JX3  A  dX4  A  rfxi  A  rfx3  A  dx$ 

+  X^X6  Jxj  A  dX2  A  dX2  A  dX4  A  dxg 

+  (2X3  —  X2)X(,  dx\  A  JX3  A  dX2  A  C/X4  A  dx(, 

+  eA3X6  JX3  A  dX4  A  C/X2  A  C/X4  A  dx§. 


8.1  |  Exercises  535 


Because  of  formula  (4),  most  of  the  terms  in  this  sum  are  zero.  In  fact, 

co  A  ?7  =  (2x3  —  X2)x(>  dx\  A  dxi,  A  dxi  A  dx4  A  dxc, 
=  (X2  —  2xt, )X6  dx\  A  dX2  A  dx?,  A  dx4  A  dX(,, 
using  formula  (3).  ♦ 

From  the  various  definitions  and  observations  made  so  far,  we  can  estab- 
lish the  following  results,  which  are  useful  when  computing  with  differential 
forms: 

PROPOSITION  1 .4  (Properties  of  the  exterior  product)  Assume  that  all 
the  differential  forms  that  follow  are  defined  on  U  c  R" : 

1.  Distributivity.  If  co\  and  002  are  &-forms  and  r\  is  an  /-form,  then 

(&>i  +  002)  A  Y]  =  co\  A  rj  +  0J2  A  rj. 

2.  Anticommutativity.  If  co  is  a  &-form  and  rj  an  /-form,  then 

CO  AT]  =  (—  \)H7]  A  CO. 

3.  Associativity.  If  co  is  a  k-foxm,  rj  an  /-form,  and  x  a  p-form,  then 

(a>  A  rj)  A  x  =  co  A  (rj  A  x). 

4.  Homogeneity.  If  co  is  a  /c-form,    an  /-form,  and  /  a  0-form,  then 

(fco)  A  rj  =  f(co  A  rj)  =  co  A  (frj). 


8.1  Exercises 

Determine  the  values  of  the  following  differential  forms  on  the 
ordered  sets  of  vectors  indicated  in  Exercises  1—7. 

1.  dx\  —  3dx2;  a  =  (7,  3) 

2.  2dx  +  6dy-5dz;  a  =  (1,-1,-2) 

3.  3dxi  a  dx2;  a  =  (4,  -1),  b  =  (2,  0) 

4.  Adx  Ady  -Idy  A  dz;  a  =  (0,  1, -1),  b  =  (1,  3,  2) 

5.  IdxAdyAdz;  a  =  (1,0, 3),  b  =  (2,-l,0),  c  = 
(5,2,  1) 

6.  dx\  A  dx2  +  2dx2  A  t/x3  +  3^x3  A  dx^,    a  =  (1,  2, 
3,4),  b  =  (4,  3,2,  1) 

7.  2dxi  A  d^3  A  dx4  +  dx2  A  JX3  A  dx$;       a  =  (1,  0, 
-1,  4,  2),  b  =  (0,  0,  9,  1,  -1),  c  =  (5,  0,  0,  0,  -2) 

8.  Let  w  be  the  1-form  on  R3  defined  by 

co  =  x2y  dx  +  y2z  dy  +  z3x  dz. 

Find  &)(3,_ii4)(a),  where  a  =  (a\,ai,  a{). 

9.  Let  co  be  the  2-form  on  R4  given  by 

co  =  X1X3  dx\  A  dx?,  —  X2X4  dx2  A  dx4. 
Find  w(2,_i,_3,i)(a,  b). 


10.  Let  co  be  the  2-form  on  R3  given  by 

co  =  cosxJxA  dy  —  sinz  dyA  dz  +  (y2  +  3)dxA  dz. 
Find  <W(o,-i,jr/2)(a>  b),  where  a  =  (ai,a2,aj,)  and 

b  =  (bi,b2X). 

11.  Let  co  be  as  in  Exercise  10.  Find  W(t  y  z)((2,  0,  —1), 
(1,7,5)). 

1 2.  Let  co  be  the  3-form  on  R3  given  by 

co  =  (ex  cos  y  +  (y2  +  2)e2z)  dx  A  dy  A  dz. 

Find  o>(o,o,o)(a,  b,  c),  where  a  =  (ai,  a.2, 03),  b  = 
(bub2,  bi),  and  c  =  (ci,  c2,  c3). 

13.  Let  a;  be  as  in  Exercise  12.  Find  co(x  y  ^((\,Q,Q), 
(0,  2,  0),  (0,  0,  3)). 

In  Exercises  14—19,  determine  co  A  rj. 

14.  On     R3:     co  =  3  dx  +  2dy  —  x  dz;     rj  =  x2dx — 
cosydy  + 1  dz. 

1 5.  On  R3 :  co  =  y  dx  —  x  dy;  rj  =  z  dx  A  dy  +  y  dx  A 
dz  +  x  dy  A  dz. 


536 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


16. 


17. 


21.  Prove  formula  (4).  (Hint:  Use  formula  (3).) 


22.  Explain  why  a  fc-form  on  R"  with  k  >  n  must  be  iden- 
tically zero. 


23.  Prove  property  1  of  Proposition  1.4. 


18. 


On  R4:  co  =  (x\  +  X2)dx\  A  dx%  A  dxi  +  (x-j  —  X4) 
dx\  A  dx2  A  dxi\\  rj  =  x\  dx\  +  2x2  dx2  +  3x3  dx$. 


24.  Prove  property  2  of  Proposition  1 .4.  (Hint:  Use  for- 


mula (3).) 


19. 


On  R5:  co  =  x\  dxi  A  dx^  —  X2X3  dx\  A  dx$; 

rj  =  eXiXidx\  A  dx4  A  dx^  —  X\  COSX5  dx2  A  dx^  A 

dx4. 


25.  Prove  property  3  of  Proposition  1.4. 

26.  Prove  property  4  of  Proposition  1.4. 


20. 


Prove  formula  (3)  by  evaluating  dxi{  A  dxi2  A  •  •  •  A 
dxik  on  k  vectors  ai, . . . ,  a*  in  R". 


8.2  Manifolds  and  Integrals  of  k -forms 

In  this  section,  we  investigate  how  to  integrate  &-forms  over  ^-dimensional  objects 
(i.e.,  curves,  surfaces,  and  higher-dimensional  analogues)  in  R". 

Integrals  over  Curves  and  Surfaces   

We  begin  by  considering  integrals  of  1 -forms  and  2-forms  over  parametrized 
curves  and  surfaces. 


DEFINITION  2.1  Let  x:  [a,  b]  R"  be  a  C1  path  in  R".  If  co  is  a  1-form 
defined  on  an  open  set  U  c  R"  that  contains  the  image  of  x,  then  the  integral 
of  co  over  x,  denoted  fx  co,  is 


EXAMPLE  1  Let  co  =  (x2  +  y)  dx  +  yz  dy  +  (x  +  y  -  z)  dz.  We  integrate  co 
over  the  path  x:  [0,  1]      R3,  x(t)  =  (2t  +  3,  3r,  7  -  t). 

We  have  x'(t)  =  (2,  3,  —  1)  so  that,  using  Definition  2.1,  we  find  that 


0=1    a>(2t+3,3t,7-o(2»  3, -l)dt 
Jo 


+  (It  +  3  +  3t  -  (7  -  t))dz(2,  3,-1)]  rfr 


♦ 


In  general,  if 


co  =  F\  dx\  +  F2  dx2  +  h  Fn  dx, 


8.2  |  Manifolds  and  Integrals  of  k-forms  537 


is  a  1-form  on  R"  and  x:  [a,  b]  — >  R"  is  any  path,  then 


«xW(x'(0)  =  F1(x(0)^1(x'(0)  +  F2(x(t))dx2(x'(t)) 
+  ■■■  +  Fn(x(t))dxn(x'(t)) 
=  FiixQMit)  +  F2(x(/))4(0  +  ■  ■  ■  +  F„(x(0K(f) 
=  (F1(x(0),F2(x(0),...,Fn(x(0))-x'(0. 
From  this  we  conclude  the  following: 

PROPOSITION  2.2  If  F  denotes  the  vector  field  (FltF2,...,  Fn)  and 

co  =  F\  dx\  +  F2  dx2  -\  +  Fn  dx„ 

and  if  x:  [a,  b]  — >  R"  is  a  C1  path,  then 

fco=f  cox(n(x'(t))dt  =  f  F(x(t))-x'(t)dt  =  f  F-ds. 


That  is,  integrating  a  1-form  over  a  path  (or,  indeed,  over  a  simple,  piecewise  C 
curve)  is  exactly  the  same  as  computing  a  vector  line  integral. 


Now  we  see  how  to  integrate  2-forms  over  parametrized  surfaces  in  R3 . 


DEFINITION  2.3  Let  D  be  a  bounded,  connected  region  in  R2  and  let 
X:  D  — »  R3  be  a  smooth  parametrized  surface  in  R3.  If  co  is  a  2-form  defined 
on  an  open  set  in  R3  that  contains  X(D),  then  we  define  fx  on,  the  integral 
of  co  over  X,  as 


(Recall  that  T.v  =  3X/3<f  and  T,  =  dX/dt.) 


Let's  work  out  the  integral  in  Definition  2.3.  We  write  co  as 

F\  dy  Adz  +  F2  dz  A  dx  +  F3  dx  A  dy 
and  X(s,  /)  as  (x(s,  t),  y(s,  t),  z(s,  t)).  Therefore, 


1 1  [Fi(X(s,t))dy  Adz(Ts,T,)+  F2(X(s,t))dz  Adx(Ts,Tt) 


J  JD 


+  F3(X(s,  t))  dx  A  dy(Ts ,  Tr)]  ds  dt. 
By  definition  of  the  basic  2-forms, 


dy  A  Jz(TJ;T,)  =  det 


dy(Ts) 
dz(Ts) 


dy(Tt) 
dz(Tt) 


=  det 


dy/ds 
dz/ds 


dy/dt 
dz/dt 


3(y,z) 
d(s,  t) 


Chapter  8 


Vector  Analysis  in  Higher  Dimensions 
Similarly,  we  have 


3(z,  x)  d(x,  y) 

dz  A  dx(Ts,  Tf)  =   and    dx  A  dy(Ts,  T,)  = 


d(s,t) 

Hence,  if  F  =  (F1;  F2,  F3),  then 


3(J,  0 


W/.[ 


Fi(X(jr,  0)^7  r  +  F2(X(s,  t))- 


+  F3(X(s,t)) 


d(s,  t) 
d(x,  y) 


d(s,  t) 


9(5,  0 


ds  dt 


-a 


D  V  3(M)    3(M)  9(M). 


Recall  from  formula  (7)  in  §7.1  that 

3(y,  z)  3(z,  ■*)  9(*.  y) 
3(s,0'  3(5,  0  '  3(s,  0 


N(s,  0, 


the  normal  to  X(Z))  at  the  point  X(s,  f).  Therefore,  we  have  established  the  fol- 
lowing Proposition  (see  also  Definition  2.2  of  Chapter  7): 

PROPOSITION  2.4  If  F  denotes  the  vector  field  F  =  Fj  i  +  F2  j  +  F3  k  and 

a)  =  F\  dy  A  dz  +  F2  dz  /\  dx  +  F3      A  dy 

and  if  X:  £)  — »  R3  is  a  smooth  parametrized  surface  such  that  co  (or  F)  is  defined 
on  an  open  set  containing  X(D),  then 


L--II. 


<WxCj,o(Ts,  T,)dsdt 


II. 

-IL 


F(X(s,  0)-N(s,  0^ 


Parametrized  Manifolds   

Next,  we  generalize  the  notions  of  parametrized  curves  and  surfaces  to 
higher-dimensional  objects  in  R".  To  set  notation,  let  R^  have  coordinates 

(U\,  U2  Wjfc). 


DEFINITION  2.5  Let  D  be  a  region  in  RA  that  consists  of  an  open,  connected 
set,  possibly  together  with  some  or  all  of  its  boundary  points.  A  parametrized 
A-manifold  in  R"  is  a  continuous  map  X:  D  — >•  R"  that  is  one-one  except, 
possibly,  along  3D.  We  refer  to  the  image  M  =  X(D)  as  the  underlying 
manifold  of  X  (or  the  manifold  parametrized  by  X). 

Such  a  & -manifold  possesses  k  coordinate  curves  defined  from  X  by 
holding  all  the  variables  U  \ ,  .  .  .  ,  Ufa  fixed  except  one;  namely,  the  jth  coor- 
dinate curve  is  the  curve  parametrized  by 

uj  1 — >  X(a\, . . . ,  dj-i,  Uj,  Uj+\, . . . ,  at), 


8.2  |  Manifolds  and  Integrals  of  k-forms  539 


where  the  a,  's  (i  ^  j )  are  fixed  constants.  If  X  is  differentiable  and  x\ ,  X2  

xn  denote  the  component  functions  of  X,  then  the  tangent  vector  to  the  jth 
coordinate  curve,  denoted  TM  ,  is 

3X      /  dx\    3x2  9x„ 


j  = 

'      duj      \duj '  duj '      '  duj 


A  parametrized  k -manifold  is  said  to  be  smooth  at  a  point  X(uo)  if  the 
mapping  X  is  of  class  C1  in  a  neighborhood  of  Uo  and  if  the  k  tangent  vectors 
TMl , . . . ,  TUl  are  linearly  independent  at  X(uo).  (Recall  that  k  vectors  vi , . . . ,  v& 
inR"  are  linearly  independent  if  the  equation  civi  +  •  •  •  +  c^k  =  0  holds  if  and 
only  if  c\  =  c2  =  ■  ■  ■  =  ck  =  0.)  A  parametrized  k -manifold  is  said  to  be  smooth 
if  it  is  smooth  at  every  point  of  X(uo)  with  uo  in  the  interior  of  D. 

Sometimes  we  will  refer  to  the  underlying  manifold  M  =  X(D)  of  a  para- 
metrized manifold  X:  D  — >  R"  as  a  parametrized  manifold;  we  do  not  expect  any 
confusion  will  result  from  this  abuse  of  terminology. 

EXAMPLE  2   Let  D  =  [0,  1]  x  [1,  2]  x  [-1,  1]  andX:  D  -+  R5  be  given  by 

X(wi,  U2,  W3)  =  (Ml  +  U2,  3«2>  W2W3,  U2  —  M3,  5^3). 

We  show  that  M  =  X(D)  is  a  smooth  parametrized  3-manifold  in  R5. 

Note  first  that  X  is  continuous  (in  fact,  of  class  C°°)  since  its  component 
functions  are  polynomials.  To  see  that  X  is  one-one,  consider  the  equation 

X(u)  =  X(u);  (1) 

we  show  that  u  =  u.  Equation  (1)  is  equivalent  to  a  system  of  five  equations: 

U\  +  U-2  =  U\  +  U2 
3U2  =  3^2 
W2«3  =  M2M3 
U2  —  W3  =  U2  —  W3 
5«3  =  5^3 

The  second  equation  implies  u2  =  U2,  and  the  last  equation  implies  u3  =  S3. 
Hence,  the  first  equation  becomes 

U\+U2  =  U\+U2       <S=>  U\=U\. 

Thus, 

u  =  {u\,  U2,  M3)  =  (Hi,  ui,  U3)  =  u. 

To  check  the  smoothness  of  M,  note  that  the  tangent  vectors  to  the  three 
coordinate  curves  are 

TUI  =  —  =(1,0,  0,0,0); 
au\ 

9X 

T„2  =  —  =  (1,3,4  1.0); 
ax 

T„3  =  - —  =  (0,  0,  2w2m3,  -1,  5). 
d«3 


540       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Therefore,  to  have  ciTi  +  C2T2  +  C3T3  =  0,  we  must  have 

(ci  +  C2,  3c2,  u\c2  +  2M2M3C3,  C2  —  C3,  5c3)  =  (0,  0,  0,  0,  0). 

It  readily  follows  that  c\  =  c2  =  C3  =  0  is  the  only  possibility  for  a  solution. 
Hence,  T„, ,  T„2,  T„3  are  linearly  independent  at  all  u  e  D  and  so  M  is  smooth  at 
all  points.  ♦ 


Figure  8.1  The  planar  robot  arm 
of  Example  3 .  Each  rod  is  free  to 
pivot  about  the  appropriate  linkage 
points. 


Parametrized  k -manifolds,  although  seemingly  abstract  mathematical  notions 
when  k  is  larger  than  3,  are  actually  very  useful  for  describing  a  variety  of  situa- 
tions, one  of  which  is  illustrated  in  the  next  example. 

EXAMPLE  3  A  planar  robot  arm  is  constructed  consisting  of  three  linked  rods 
of  lengths  1,  2,  and  3.  (See  Figure  8.1.)  The  rod  of  length  3  is  anchored  at  the 
origin  of  R2  but  free  to  rotate  about  the  origin.  The  rod  of  length  2  is  attached  to 
the  free  end  of  the  rod  of  length  3,  and  the  rod  of  length  1  is,  in  turn,  attached  to 
the  free  end  of  the  rod  of  length  2.  We  describe  the  set  of  positions  that  the  arm 
can  take  as  a  parametrized  manifold. 

Clearly,  each  state  of  the  robot  arm  is  determined  by  the  coordinates  (x\ ,  yi ), 
(*2 ,  yi),  and  (x3 ,  y3)  of  the  linkage  points,  which  we  may  consider  to  form  a  vector 
x  =  (jci,  y\,  x2,  y2,  *3,  V3)  in  R6.  However,  not  all  vectors  in  R6  represent  a  state 
of  the  robot  arm.  In  particular,  the  point  (x\ ,  yi)  must  lie  on  the  circle  of  radius  3, 
centered  at  the  origin,  the  point  (x2 ,  y2)  must  lie  on  the  circle  of  radius  2,  centered 
at  (jti,  yi),  and  the  point  (x3,  y3)  must  lie  on  the  circle  of  radius  1,  centered  at 
(x2,  y2).  Thus,  for  x  =  (jti,  y\,  x2,  yi,  yi)  to  represent  a  state  of  the  robot  arm, 
we  require 


x\  +  y\  =  9 

(x2  -  Xl)2  +  (y2  -  yi)2 

(x3  -  X2f  +  (V3  -  V2)2 


4. 
1 


(2) 


and 


We  may  parametrize  each  of  the  circles  in  the  system  (2)  in  a  one-one  fashion  by 
using  three  different  angles  9\,  02,  and  ©3.  Hence,  we  find 

(jci,  yi)  =  (3  cos 0i,  3  sin0i), 

(x2,  y2)  =  (x\  +  2  cos 02,  yi  +  2  sin02) 

=  (3  cos  0i  +  2  cos  02,  3  sin 0!  +2sin02),  (3) 

(*3»  y3)  =  O2  +  cos 03,  y2  +  sin03) 

=  (3  cos  0i+2  cos  02  +  cos  03 ,  3  sin  0i  +  2  sin 02  +  sin 03), 

whereO  <  0i,  02  ,  03  <  27T .  Therefore,  the  map  X:  [0,  27r)  x  [0,  27i)  x  [0,  27r)  — > 
R6  given  by 

X(0i,  02  ,  03)  =  (xi,yi,x2,  y2,  x3,  y3), 

where  (xi,  y\,  x2,  y2,  x3,  y3)  are  given  in  terms  of  0i,  02,  and  03  by  means  of  the 
equations  in  (3),  exhibits  the  set  of  states  of  the  robot  arm  as  a  parametrized  3- 
manifold  in  R6.  We  leave  it  to  you  to  check  that  X  defines  a  smooth  parametrized 
3 -manifold.  ♦ 


8.2  |  Manifolds  and  Integrals  of  k-forms  541 


Just  like  a  parametrized  surface,  a  parametrized  ^-manifold  M  =  X(D)  may 
or  may  not  have  a  boundary,  denoted  as  usual,  by  dM.  If  M  has  a  nonempty 
boundary,  then  dM  is  contained  in  the  image  under  X  of  the  portion  of  the 
boundary  of  the  domain  region  D  that  is  also  part  of  D.  Under  suitable  (and 
mild)  hypotheses,  dM,  if  nonempty,  is,  in  turn,  a  union  of  finitely  many  (k  —  1)- 
manifolds  (without  boundaries). 

EXAMPLE  4   Let  B  c  R3  denote  the  closed  unit  ball  {u  =  (hj ,  w2,  h3)  |  u\  + 

u\  +  u\<  1},  and  define  X:  B  ->  R4  by 


Then  M  =  X(B)  is  a  portion  of  a  "generalized  paraboloid"  having  equation  w  = 
x2  +  y2  +  z2;  we  have  M  =  {(x,  y,  z,  w)  £  R4  I  w  =  x2  +  y2  +  z2,  x2  +  y2  + 
z2  <  1}.  In  this  case,  dM  =  [(x,  y,  z,  1)  |  x2  +  y2  +  z2  =  1}.  Note  that  dM  is  a 
parametrized  2-manifold  in  R4,  as  we  may  see  via  the  map 

Y:  [0, 7r]  x  [0,  2tt)  — >  R4,     Y(i,  t)  =  (sins  cost,  sins  sin?,  coss,  1).  ♦ 

Integrals  over  Parametrized  ^-manifolds  — 

Now,  we  see  how  to  define  the  integral  of  a  k-ioxm  over  a  smooth  parametrized 
A- -manifold.  Our  definition  generalizes  those  of  Definitions  2.1  and  2.3. 


DEFINITION  2.6    Let  D  be  a  bounded  connected  region  in  R*  and  X:  D 
R"  a  smooth  parametrized  ^-manifold.  If    is  a  &-form  defined  on  an  open 
set  in  R"  that  contains  M  =  X(D),  then  we  define  the  integral  of  00  over  M 
(denoted fx  co)  by 


(Here  f  ■  ■  ■  f  refers  to  the  ^-dimensional  integral  over  D.) 


EXAMPLE  5  Let  X:  [0,  1]  x  [1,  2]  x  [-1,  1]  ->  R5  be  the  parametrized  3- 
manifold  defined  by 


(See  Example  2.)  Let  co  be  the  3 -form  defined  on  R5  as 

co  =  X1X3  dx\  A  dxj,  A  dxs  +  (X3X4  —  2x2x5)  dx2  A  dx4  A  JX5. 
We  calculate  fx  co. 

Recall  from  Example  2  that  the  tangent  vectors  to  the  three  coordinate  curves 

are 


X(«i,  M2,  W3)  =  (hi,  H2,  H3,  u\  +  u\  +  h2). 


X(Hi,  H2,  M3)  =  (Hi  +  H2,  3H2,  M2H3,  H2  —  H3,  5H3). 


T 


T 


(1,0,  0,  0,  0), 
(1,3,h2,  1,0), 


and 


T, 


(0,  0,2h2h3, -1,5). 


542       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Then,  from  Definition  2.6, 


L 


=  +  u2)u2u\dxi  Adx3  Adxs(Tul,  TU2,  TM3) 

J-\  Ji  Jo 

+  (u2u\(u2  —  M3)  —  30^2^3)^2  A  ^4  A  ^X5(TM1,  T„2,  T„3)}  du\  du2  duT, 


f-i  Ii  lo 


(ui  +  u2)u2u3 


1  1 


0        W3  2m2"3 


0  0 


0 

3 

0 

+  (M2M3  —  W2M3  —  3OM2M3) 

0 

1 

-1 

0 

0 

5 

du\  du,2  duT, 


/I  I  5(ui  +  U2)U2u\  du\  du2  du-i, 
1  J\  Jo 


37 
6  ■ 


EXAMPLE  6   If  <w  is  a  3-form  on  R3 ,  then  co  may  be  written  as 

co  =  F(x,  y,  z)dx  A  dy  A  dz. 

(Why?)  If  D*  is  a  bounded  region  in  R3  and  X:  D*  ->  R3  is  a  smooth  parametrized 
3-manifold,  then  Definition  2.6  tells  us  that 

j  0}  =  J J  j  COx(uuui:u3)(J'ui,Tu2>r^u3)dUldu2dU3 

=  III  F(X(u))dx  A  dy  A  dz(TUl,  TM2,  T„3)<iMi  du2  dui 


dx/dui  dx/du2  dx/dui, 
dy/dui  dy/du2  dy/dui, 
dz/dui       dz/du2  3z/3w3 


du\  du2  du-i 


d(x,  y,  z) 

F(x(u),  y(u),  z(u))—   du\ du2  du3 

D*  d{uUU2,Uyj 


III 

±  j  j  j  F(x,  y,  z)dx  dy  dz, 


from  the  change  of  variables  theorem  for  triple  integrals  (Theorem  5.5  of  Chapter 
5),  where  D  =  X(D*).  ♦ 

Orientation  of  a  Parametrized  Zj-manifold   

We  have  seen  that  vector  line  integrals  and  vector  surface  integrals  may  be  defined, 
respectively,  over  oriented  curves  and  surfaces  in  a  manner  effectively  independent 
of  the  parametrization  used.  We  now  see  how  it  is  possible  to  define  the  integral 


8.2  |  Manifolds  and  Integrals  of  k-forms  543 


C 

Figure  8.2  An  orientation  of  the 
curve  C  shown  is  a  choice  of 
continuously  varying  unit  tangent 
vector  T  along  C. 


N  N 


Figure  8.3  An  orientation  of 
the  surface  S  is  a  choice  of 
continuously  varying  unit  normal 
vector  N  along  S. 


of  a  £-form  over  a  parametrized  & -manifold  X:  D  —>  R"  so  that  it  depends  largely 
on  the  underlying  manifold  M  =  X(D),  rather  than  on  the  particular  map  X.  To 
do  this,  we  must  consider  how  reparametrization  of  M  affects  the  integral,  and 
we  must  define  what  we  mean  by  an  orientation  of  M. 

First,  we  consider  the  notion  of  orientation.  We  have  previously  seen  how 
parametrized  curves  and  surfaces  can  be  oriented  by  using  some  fairly  natural 
geometric  ideas.  A  smooth  parametrized  curve  implicitly  received  an  orientation 
from  the  parameter;  typically,  we  orient  a  curve  by  indicating  the  direction  in 
which  the  parameter  variable  increases.  We  may  also  think  of  an  orientation  of 
a  curve  as  a  choice  of  a  unit  tangent  vector  T  at  each  point  of  the  curve,  made 
so  that  T  varies  continuously  as  we  move  along  the  curve.  (See  Figure  8.2.)  An 
orientation  of  a  smooth  parametrized  surface  in  R3,  when  it  exists,  is  a  choice 
of  a  continuously  varying  unit  normal  vector  N  at  each  point  of  the  surface.  (See 
Figure  8.3.) 

To  define  notions  of  orientation  and  orientability  for  a  parametrized  k- 
manifold  when  k  >  2,  we  will  need  to  work  more  formally. 

First,  we  need  to  introduce  two  related  ideas  from  the  linear  algebra  of  R" . 
Thus,  suppose  Vi,  \i, . . . ,  \k  are  vectors  in  R".  By  a  linear  combination  of 
Vi , . . . ,  Vfc,  we  mean  any  vector  v  e  R"  that  can  be  written  as 

v  =  civi  +c2v2H  \-ckvk 

for  suitable  choices  of  the  scalars  c\ ,  . . .  ,ck.  The  set  of  all  possible  linear  com- 
binations of  vi  V£,  called  the  (linear)  span  of  vi ,  . . . ,  \k,  will  be  denoted 

Span{vi , . . .  ,\k}.  That  is, 


Span{v! , 


{civi  H  \-ck\k  \ci,...,cke  R}. 


DEFINITION  2.7  Let  M  =  X(D),  where  X:  D  C  R*  R",  be  a  smooth 
parametrized  k -manifold.  An  orientation  of  M  is  a  choice  of  a  smooth, 
nonzero  &-form  Q  defined  on  M.  If  such  a  &-form  £1  exists,  M  is  said  to  be 
orientable  and  oriented  once  a  choice  of  such  a  A'-form  is  made. 


Although  we  cannot  readily  visualize  an  orientation  £2  of  a  parametrized  k- 
manifold  when  k  is  large,  we  can  nonetheless  see  how  the  tangent  vectors  to  the 
coordinate  curves  relate  to  it. 


DEFINITION  2.8  Let  M  =  X(D)  be  a  smooth  parametrized  k -manifold 
oriented  by  the  A-form  Q.  The  tangent  vectors  TUl , . . . ,  T„4  to  the  coordinate 
curves  of  M  are  said  to  be  compatible  with  Q  if 


^x(u)(Tul 


T„J  >  0. 


We  also  say  that  the  parametrization  X  is  compatible  with  the  orientation  Q 


if  the  corresponding  tangent  vectors  TUl 


,  T„,  are. 


Note  that  if  TMl  TUk  are  incompatible  with  the  orientation  Q,  then  they 

are  compatible  with  the  opposite  orientation  —  £1.  Alternatively,  we  may  change 
the  parametrization  X  of  M  by  reordering  the  variables  u\,  . . . ,  uk  to,  say,  112,  ui, 


«3, 


sothatT„2,TUl, 


TUk  are  compatible  with  Q.. 


544       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Definition  2.7  is  consistent  with  the  earlier  definitions  of  orientations  of 
curves  and  surfaces,  as  we  now  discuss.  Suppose  first  that  x:  /  ->  R"  is  a  smooth 
parametrized  curve  in  R"  (where  /  is  an  interval  in  R)  and  T  is  a  continuously 
varying  choice  of  unit  tangent  vector  along  C  =  x(7).  Then  we  may  define  an 
orientation  1-form  £2  on  C  by 

^x(r)(a)  =  T  •  a. 

Conversely,  given  an  orientation  1-form  Q,  we  may  define  a  continuously  varying 
unit  tangent  vector  from  it  by  taking  T  to  be  the  unique  unit  vector  parallel  to 
x'(0  such  that,  for  any  nonzero  vector  a  parallel  to  x'(?), 

T  •  a  has  the  same  sign  as  £2x(f)(a). 

That  T  is  uniquely  determined  follows  because  T  must  equal  ±xr(t)/\\x'(t)\\,  so 
knowing  a  and  the  value  of  £2x(f)(a)  determines  the  choice  of  sign  for  T. 

Similarly,  suppose  5  =  X(D)  is  a  smooth  parametrized  surface  in  R3  (i.e.,  a 
smooth  parametrized  2-manifold).  If  we  can  orient  S  by  a  continuously  varying 
unit  normal  N,  then  we  may  define  an  orientation  2-form  Q  on  5  by 

fiX(«1,«2)(a,b)  =  det[  Nab], 

where  [  N  a  b  ]  is  the  3x3  matrix  whose  columns  are,  in  order,  the  vectors 
N,  a,  b.  Conversely,  given  an  orientation  2-form  Q  on  5,  we  may  define  a  con- 
tinuously varying  unit  normal  N  from  it  by  taking  N  to  be  the  unique  unit  vector 
perpendicular  to  T„,  and  T„2  (and  hence  to  every  vector  in  Span{T„, ,  T„2})  such 
that,  for  any  pair  a,  b  of  linearly  independent  vectors  in  SpanjT,,, ,  T„2}, 

det  [  N  a  b  ]  has  the  same  sign  as  £2X(i(1,„2)(a,  b). 

To  see  that  N  is  uniquely  determined,  note  that,  given  linearly  independent  vectors 
a,  b  in  Span{TI(1 ,  T„2},  the  only  possibilities  for  N  are 

T    x  T 

I  U 1  H2 


|T„.  x  Tu 


Hence,  we  choose  the  sign  for  the  normal  vector  N  so  that  det  [Nab]  has 
the  same  sign  as  f2x(Ml,M2)(a,  b). 

EXAMPLE  7  Consider  the  generalized  paraboloid  M  =  {(x,  y,  z,  w)  e  R4  | 
w  =  x2  +  y2  +  z2},  which  we  may  exhibit  as  a  smooth  parametrized  3-manifold 
via 

X:  R3  — >  R4,     X(«i,  U2,  w3)  =  {u\,  u%,  m3,  u\  +  u\  +  u2). 

We  show  how  to  orient  M. 

Note  that  the  equation  x2  +  y2  +  z2  —  w  =  0  shows  that  M  is  the  level  set  at 
height  0  of  the  function  F(x,  y,  z,  w)  =  x2  +  y2  +  z2  —  w.  Hence,  the  gradient 
V  F  =  (2x  ,2y,2z,  —  1 )  is  a  vector  normal  to  M .  If  we  employ  the  parametrization 
X  and  normalize  the  (parametrized)  gradient,  we  see  that 

(2wi,2k2,2w3,  -1) 

N(Mi,  U2,  M3)  = 


' 4u2  +  Au\  +  Au\  +  1 
is  a  continuously  varying  unit  normal.  Moreover,  the  3-form  Q,  defined  on  M  as 
^x(u)(ai,  a2,  a3)  =  det  [N  ai   a2  a3] 


8.2  |  Manifolds  and  Integrals  of  k-forms  545 


gives  an  orientation  for  M.  Note  that 


=  det 


I  T„ 

T„    T„  1 

1    Aj,^-  _l_  ^.ij2 

T  1 

^Au\ 

+  Au\  +  Au\ 

+  1 

^Au\ 

+  Au\  + 

+  1 

-1 

^Au\ 

+  Au\  + 

+  1 

+  4u22 

+  4M2+  1. 

1      0  0 


0  0 


2wi   2u2  2«3 


Since  this  last  expression  is  strictly  positive,  we  see  that  TM1,  T„2,  TU3  are  com- 
patible with  £2.  ♦ 

EXAMPLE  8   We  may  generalize  Example  7  as  follows: 

Suppose  that  M  c  R"  is  the  graph  of  a  function  /:  U  C  R"-1  R";  that 
is,  suppose  M  is  defined  by  the  equation  jc„  =  /(xi , . . . ,  x„_i).  Then  M  may  be 
parametrized  as  an  (n  —  1  )-manifold  via 


Kn-l))- 


X:U  c  R"  1  ->  R\  X(Ml,  ...,uB_i)  =  (mi,...,  w„_i,  /(m, 
Since  Af  is  also  the  level  set  at  height  0  of  the  function 

F(x\,  . . . ,  xn)  =  f(X\, . . . ,  x„_i)  —  xn, 


a  vector  normal  to  Mis  provided  by  the  gradient  VF  =  (fXl, . . . ,  fx„_lt  —  1).  Ifwe 
normalize  VF  and  use  the  parametrization  X,  we  see  that  we  have  a  continuously 
varying  unit  normal 


N(m 


,  Un-\) 


{fill  I  •  •  •  !   fll„-l  !  1) 


from  which  we  may  define  our  orientation  (n  —  l)-form  £2  for  M  by 
fix(u)(ai,...,a„_i)  =  det[N  ai   ■•■  a„_i]. 


Now  suppose  that  M  is  a  smooth  parametrized  £ -manifold  in  R"  with  non- 
empty boundary  dM.  If  M  is  oriented  by  the  &-form  £2,  then  there  is  a  way  to 
derive  from  it  an  orientation  for  dM,  which  we  describe  in  Definition  2.9.  To 
set  notation,  let  X:  D  C  R*  — >•  R"  denote  the  parametrization  of  M  and  suppose 
Y:  E  c  RA'~'  — R"  gives  a  parametrization  of  a  connected  piece  of  3M  as  a 
smooth  (k  —  l)-manifold.  Since  dM  is  part  of  M,  if  s  =  (si , . . . ,  S£_i)  e  E,  then 
there  is  some  u  =  (u\, . . . ,  Uk)  e  D  such  that  Y(s)  =  X(u). 


546       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


DEFINITION  2.9  Let  M  be  a  smooth  parametrized  /: -manifold  in  R"  with 
boundary  dM.  Suppose  M  is  oriented  by  the  &-form  Q.  Then  the  connected 
pieces  of  dM  are  said  to  be  oriented  consistently  with  M,  or  that  dM  has 
its  orientation  induced  from  that  of  M ,  if  the  orientation  (k  —  l)-form  Q3M 
is  determined  from  Q  as  follows.  Let  V  be  the  unique,  outward-pointing  unit 
vector  in  R",  defined  and  varying  continuously  along  dM,  that  is  tangent  to 
M  and  normal  to  dM.  (See  Figure  8.4.)  Then  £2aM  is  defined  as 

n£«(«i. . . . ,  ak-i)  =  C2X(u)(V,  alt . . . ,  aw), 

where  the  map  X:  D  c  R*  ^  R"  parametrizes  M,  the  map  Y:  E  c  Rk~l  -> 
R"  parametrizes  a  connected  piece  of  dM,  and  Y(s)  =  X(u). 

Note  that,  in  particular,  the  vector  V  in  Definition  2.9  must  be  such  that 


•  V  e  Span{T„, , . . . ,  T„J  (i.e.,  V  is  tangent  to  M); 

■  V-  TSj  =  0  for  i  =  1, . . . ,  k  -  1  (i.e.,  V  is  normal  to  dM); 

'  V  points  away  from  M. 


These  conditions  are  often  not  difficult  to  achieve  in  practice.  Definition  2.9  will 
be  very  important  when  we  consider  a  generalization  of  Stokes's  theorem  in  the 
next  section. 

EXAMPLE  9    Consider  the  surface  5  in  R3  consisting  of  the  portion  of  the 
cylinder  x2  +  y2  =  4  with  2  <  z  <  5.  Note  that  the  boundary  of  S  consists  of  the 
two  circles  {(x,  y,  z)  \  x2  +  y2  =  4,  z  =  2}and{(x,  y,  z)  \  x2  +  y2  =  4,  z  =  5}. 
We  investigate  how  to  orient  35  consistently  with  an  orientation  of  S. 
The  cylinder  may  be  parametrized  as  a  2 -manifold  in  R3  by 

X:  [0,  2tt)  x  [2,  5]  — >  R3,        X(wi,  ui)  =  (2cosmi,  2sinwi,  uj). 

Then  the  tangent  vectors  to  the  coordinate  curves  are 

TMl  =  (— 2sinM!,  2coswi,  0) 

and 

TU2  =  (0,  0,  1). 

Since  S  is  a  portion  ofthe  level  set  at  height  4  of  the  function  F(x,  y,  z)  =  x2  +  y2, 
a  unit  normal  N  to  S  is  given  by 

(2x,  2y,  0)  =  tx_  y  \ 
l|VF||      J4x2  +  4y2      V2'2'  )' 

In  terms  of  the  parametrization  X,  the  normal  N  is  also  given  by 

N  =  (cos  Mi,  sin  Mi,  0). 

Then  we  may  define  an  orientation  2-form  on  S  by 

£2x(jn,«2)(ai,  a2)  =  det[N  ai  a2]. 


Figure  8.4  The  outward-pointing 
unit  vector  V  of  Definition  2.9. 


8.2  |  Manifolds  and  Integrals  of  k-forms  547 


Hence, 


Figure  8.5  Orienting  the 
boundary  of  the  surface  S  of 
Example  9.  Note  the 
outward-pointing  tangent  vectors 
Vtop  and  Vbottom- 


(THl,  TH2)  =  det 


COSM] 

0 


—2  siriM] 
2  cos  u\ 
0 


2  >  0. 


Thus,  TH1,  TH2  are  compatible  with  £2. 

We  may  parametrize  35  by  using  two  mappings: 


and 


Bottom  circle:     Yi :  [0,  2n)  ->  R3 


Top  circle:     Y2:  [0,  2tt)  ->  R3 


Yi(s)  =  (2  cos 5,  2  sins,  2) 


Y2(s)  =  (2  coss,  2  sins,  5) 


To  use  Definition  2.9  to  orient  35,  we  must  identify  outward-pointing  vectors  tan- 
gent to  5  and  normal  to  35.  From  Figure  8.5,  we  see  that  along  the  top  circle  V  = 
Vtop  =  (0,  0,  1)  works,  while  along  the  bottom  circle,  V  =  Vbottom  =  (0,0,-1) 
suffices.  Hence,  Definition  2.9  tells  us  that,  along  the  bottom  circle, 

nYi(s)(a)  =  ^X(i,2)(Vbottom,  a)  =  det  [  N    Vbottom     3  ]  , 

while  along  the  top  circle, 

^)(a)  =  Qx(,,5)(Vtop,a)  =  det[N  Vtop  a]. 

For  both  maps  Yi  and  Y2,  we  have  that  the  coordinate  tangent  vector  is  Ts  = 
(—2  sins,  2  cos  s,  0).  Thus,  along  the  bottom  circle, 


Q 


dS 

YiCs) 


(Ts)  =  det 


coss 
sins 
0 


0 
0 

-1 


—2  sins 
2  cos  s 
0 


=  2, 


so  Ts  is  compatible  with  the  orientation  1-form  £ldS.  However,  along  the  top 
circle, 

~  coss   0  —2 sins 
^(s)(Tv)  =  det     sins    0  2coss 
0     1  0 


so  Ts  is  incompatible  with  £2  .  Therefore,  we  must  orient  the  top  circle  clockwise 
around  the  z-axis  and  the  bottom  circle  counterclockwise.  ♦ 


The  following  example  is  the  three-dimensional  analogue  of  Example  9: 

EXAMPLE  10  Consider  the  subset  M  c  R4givenbyM  =  {(x,  y,  z,  w)  \  x2  + 
y2  +  z2  =  4,  2  <  w  <  5}.  This  set  M  is  a  portion  of  the  cylinder  over  a  sphere 
of  radius  2.  Note  that  the  boundary  of  M  consists  of  the  two  spheres  5bottom  = 
{(*,  y,  z,  2)\x2  +  y2  +  z2  =  4}  and5top  =  {(*,  y,  z,  5)  |  x2  +  y2  +  z2  =  4}.  We 
investigate  M  and  dM  as  parametrized  manifolds,  orient  M,  and  study  the  induced 
orientation  on  dM. 

First,  we  note  that  M  may  be  parametrized  as  a  3 -manifold  in  R4  by 

X:  [0,  it]  x  [0,  2tt)  x  [2,  5]  ->  R4, 

X(«i,  «2,  K3)  =  (2  sin  mi  cos  U2,  2sinwi  sinM2,  2cosmi,  M3). 

(This  is  the  usual  parametrization  of  a  sphere  using  spherical  coordinates  (p  =  u\, 
9  =  112,  with  an  additional  parameter  M3  for  the  "vertical"  w-axis.)  The  tangent 


548       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


vectors  to  the  coordinate  curves  are  given  by 

TUl  =  (2cos«i  cos U2,  2cosmi  sinM2,  — 2sinwi,  0), 
T„2  =  (—2  sinw!  sinM2,  2  sinwi  cos  112,  0,  0), 

and 

TH3  =  (0,  0,0,  1). 

Note  that  this  parametrization  fails  to  be  smooth  when  u\  is  0  or  jt,  since  then 
T„2  =  0  at  those  values  for  u  \ .  You  can  check  that  the  parametrization  is  smooth 
at  all  other  values  of  X(u)  (i.e.,  for  u  in  (0,  jt)  x  [0,  2jt)  x  [2,  5]). 

Because  M  is  a  portion  of  the  level  set  at  height  4  of  the  function  F(x,  y, 
z,  w)  =  x2  +  y2  +  z2,  a  unit  normal  N  to  M  is  given  by 

VF  (2x,2y,2z,  0)        (x  y  z 


l|VF||      J4x2  +  4y2  +  4z2     V2' 2' 2" 
In  terms  of  the  parametrization  X,  the  normal  N  is  also  given  by 

N  =  (sinwi  cosw2,  sinwi  sinw2,  cosmi,  0). 
We  define  an  orientation  3 -form  Q  for  M  by 

^x(u)(ai,a2,a3)  =  det[N  ai   a2  a3]. 


Then 


^x(u)(TM1,  T„2,  T„3)  =  det 


sin  u\  cos  U2  2  cos  Mi  cos  i<2  — 2  sin  Mi  sin  M2  0 

sinwisinw2  2cosMisinM2  2sinwicos«2  0 

coswi  —2  sin  Mi               0  0 

0  0                    0  1 


=  4  sin  u  i  >  0 

for  0  <  wi  <  jt  (which  is  where  the  parametrization  X  is  smooth).  Hence,  THI, 
T„2,  T„3  are  compatible  with  Q.. 

We  parametrize  the  two  pieces  of  dM  with  two  mappings: 

"Bottom"  sphere  Sbottom: 

Yi:[0,tt]  x  [0,2tt)^  R4, 

Yi(^i,  52)  =  (2  sin^i  cos^2>  2sinsi  sin52,  2cos5i,  2), 

and 

"Top"  sphere  Stop: 

Y2:[0,  jt]  x  [0,2;r)^  R4, 

Y2(*i,  S2)  =  (2 sinsi  COSS2,  2sinsi  sin  52,  2cossi,  5). 

Note  that  both  parametrizations  Yi  and  Y2  give  the  same  tangent  vectors  to  the 
corresponding  coordinate  curves,  namely, 

Tj,  =  (2  cos  si  cos  52,  2  cos  si  sin  52,  —2  sinsi,  0) 

and 

TS2  =  (— 2sin5!  sin52,  2sin5!  cos 52,  0,  0), 

and  by  considering  these  tangent  vectors,  we  see  that  the  parametrizations  are 
smooth  whenever  51  ^  0,  jt. 


8.2  |  Manifolds  and  Integrals  of  k-forms  549 


To  give  dM  the  orientation  induced  from  that  of  M,  we  identify  outward- 
pointing  unit  vectors  tangent  to  M  and  normal  to  dM.  Thus,  we  need  V  such  that 

•  Ve  Span{TM1,TM2,TM3}     <=>     V-N  =  0; 

•  V-T.S1  =V-T.S2  =  0; 

•  V  points  away  from  M. 

It's  not  difficult  to  see  that  we  must  take  V  =  Vtop  =  (0,  0,  0,  1)  along  5top  and 
V  =  Vbottom  =  (0,  0,  0,  —1)  along  Sbottom-  Therefore,  Definition  2.9  tells  us  that 
along  Sbottom, 

fi^s)(ai,  a2)  =  ^x(s,2)(Vbottom,  ai,  a2). 

In  particular, 
^s)(T.s.1,TJ2)  =  det[N  Vbottom  T„  TJ2] 


=  det 


sin  5i  cos  52    0    2  cos    cos  ^2  — 2  sin  ji  sin  52 

sin^isin^    0     2cossisini2  2sinsicoss2 
cos^i        0       — 2sin^i  0 
0-10  0 


=  4  sin^i  >  0 

for  0  <  s\  <  7T  (i.e.,  where  the  parametrization  Yj  is  smooth).  Thus,  Y\  is  com- 
patible with  QdM .  Along  Stop,  however,  we  have 

^S)(T,1,TJ2)  =  det[N  Vbottom  Tn  TJ2] 


det 


sin  s\  cos  52  0  2  cos  5j  cos  52  — 2  sin  5i  sin  52 

sin 5i  sin 52  0   2cos5ishi52  2sin5iCOS52 

cos  5i  0      —2  sin  5i  0 

0  10  0 


=  — 4sin5i  <  0 

for  0  <  5i  <  it,  so  Y2  is  incompatible  with  Q3M .  We  must  take  care  with  this 
distinction  when  we  consider  the  general  version  of  Stokes's  theorem.  ♦ 

Next,  we  examine  how  the  integral  of  a  &-form  co  can  vary  when  taken 
over  two  different  parametrizations  X:  D\  —>  R"  and  Y:  D2  —>  R"  for  the  same 
^-manifold  M  =  X(D0  =  Y(D2). 


DEFINITION  2.10  Let  X:  Di  c  R1  ->  R"  and  Y:  D2cR'^  R"  be 
parametrized  k -manifolds.  We  say  that  Y  is  a  reparametrization  of  X  if  there 
is  a  one-one  and  onto  function  H:  D2  —>  D\  with  inverse  H  1  D\  — >  D2  such 
that  Y(s)  =  X(H(s)),  that  is,  such  that  Y  =  X  o  H.  If  X  and  Y  are  smooth  and 
H  and  H~ 1  are  both  class  C 1 ,  then  we  say  that  Y  is  a  smooth  reparametrization 
ofX. 


Since  H  is  one-one,  it  can  be  shown  that  the  Jacobian  det  DH  cannot  change 
sign  from  positive  to  negative  (or  vice  versa).  Thus,  we  say  that  both  H  and  Y  are 
orientation-preserving  if  the  Jacobian  det  DH  is  positive,  orientation-reversing 
if  det  DH  is  negative. 


550       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 

The  following  result  is  a  generalization  of  Theorem  2.5  of  Chapter  7  to  the 
case  of  k -manifolds. 


THEOREM  2.1 1  Let  X:  D,cR'^  R"  be  a  smooth  parametrized  k -manifold 
and  co  a  k-form  defined  on  X(£>i).  If  Y:  Di  c  —>  R"  is  any  smooth 
reparametrization  of  X,  then  either 


if  Y  is  orientation-preserving,  or 


if  Y  is  orientation-reversing. 


In  view  of  Theorem  2.1 1,  we  can  define  what  we  mean  by  fM  co,  where  M 
is  a  subset  of  R"  that  can  be  parametrized  as  an  oriented  ^-manifold  and  is  a 
/c-form  defined  on  M.  We  simply  let 


where  X:  D  C  R*  ->  R"  is  any  smooth  parametrization  of  M  that  is  compatible 
with  the  orientation  chosen. 


EXAMPLE  1 1  We  evaluate  fc  co,  where  C  is  the  (oriented)  line  segment  in  R3 
from  (0,  —  1 ,  —2)  to  (1 ,  2,  3)  and  co  =  z  dx  +  x  dy  +  y  dz. 

Using  Theorem  2.1 1,  we  may  parametrize  C  in  any  way  that  preserves  the 
orientation.  Thus, 

x:  [0,  1]  ->  R3,    x(0  =  (1  -  /)(0,  -1,  -2)  +  f(l,  2,  3)  =  (/,  3t  -  1,  5t  -  2) 

is  one  way  to  make  such  a  parametrization.  Then  x'(t)  =  (1,  3,  5)  and  hence, 
from  Definition  2.1,  we  have 


co=     co=  CL>x(t)(x'(t))dt 

JC  Jx  JO 


=  f  {(5t  -  2)  •  1  +  t  •  3  +  (3t  -  1)  •  5}  dt 
Jo 


=  f  (23t-T)dt 
Jo 


23 


It 


Note  that  if  we  parametrize  C  in  the  opposite  direction  by  using,  for  example, 
the  map 


y:  [0,  1]  ->  R3,  y(f)  =  f(0,  -1,  -2)  +  (1  -  0(1,  2, 3)  =  (1  -  f,  2  -  3r,  3  -  50, 


8.2  |  Exercises  551 


then  we  would  have 

.  i 


f  m=  f  a>y(t){y'(t))dt 

Jy  JO 

=  f  {(3  -  5t)(- 1)  +  (1  -  0(~3)  +  (2  -  3f)(-5)}  dt 
Jo 

=  jf  (23t  -  I6)dt  =  (^-t2  -  I6t 


In  light  of  Theorem  2.11,  this  result  could  have  been  anticipated  from  our  pre- 
ceding calculation  of  fx  co.  ♦ 

Note  on  k -manifolds 

The  central  geometric  object  of  study  in  this  section,  namely,  a  parametrized 
£ -manifold  is  actually  a  rather  special  case  of  a  more  general  notion  of  a  k- 
manifold.  In  general,  a  A-manifold  in  R"  is  a  connected  subset  M  c  R"  such  that, 
for  every  point  x  e  M,  there  is  an  open  set  U  c  R*  and  a  continuous,  one-one 
map  X:  U  — >  R"  with  x  e  X(t/)  c  M.  (A  ^-manifold  with  nonempty  boundary 
requires  a  somewhat  modified  definition.)  That  is,  M  is  a  (general)  k -manifold 
if  it  is  locally  a  parametrized  & -manifold  near  each  point.  It  is  possible  to  extend 
notions  of  orientation  and  integration  of  £-forms  to  this  more  general  setting, 
although  it  requires  some  finesse  to  do  so.  For  the  types  of  examples  we  are 
encountering,  however,  our  more  restrictive  definitions  suffice. 


8.2  Exercises 


1 .  Check  that  the  parametrized  3 -manifold  in  Example  3  is  in 
fact  a  smooth  parametrized  3 -manifold. 

2.  A  planar  robot  arm  is  constructed  by  using  two  rods  as 
shown  in  Figure  8.6.  Suppose  that  each  of  the  two  rods  may 
telescope,  that  is,  that  their  respective  lengths  l\  and  h  may 
vary  between  1  and  3  units.  Show  that  the  set  of  states  of 
this  robot  arm  may  be  described  by  a  smooth  parametrized 
4-manifold  in  R4. 


Figure  8.6  Figure  for  Exercise  2. 


3.  A  planar  robot  arm  is  constructed  by  using  a  rod  of  length 
3  anchored  at  the  origin  and  two  telescoping  rods  whose 
respective  lengths  h  and  /3  may  vary  between  1  and  2  units 
as  shown  in  Figure  8.7.  Show  that  the  set  of  states  of  this 


robot  arm  may  be  described  by  a  smooth  parametrized  5- 
manifold  in  R6.  (See  Exercise  2.) 


Figure  8.7  Figure  for  Exercise  3. 


4.  A  robot  arm  is  constructed  in  R3  by  anchoring  a  rod 
of  length  2  to  the  origin  (using  a  ball  joint  so  that 
the  rod  may  swivel  freely)  and  attaching  to  the  free 
end  of  the  rod  another  rod  of  length  1  (which  may 
also  swivel  freely;  see  Figure  8.8).  Show  that  the  set  of 
states  of  this  robot  arm  may  be  described  by  a  smooth 
parametrized  4-manifold  in  R6. 

5.  Suppose  vi , . . . ,  V£  are  vectors  in  R" .  If  x  e  R"  is  orthog- 
onal to  V;  for  i  =  1 , . . . ,  k,  show  that  x  is  also  orthogonal 
to  any  vector  in  Spanfvi , . . . ,  v*}. 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


-  (x2,  y2,  z2) 

xi,yi,  zi) 


Figure  8.8  Figure  for  Exercise  4. 

6.  Let  a,  and  c  be  positive  constants  and  x:  [0,  it]  — >  R3 
the  smooth  path  given  by  x(f )  =  (a  cost,  b  sin  /,  ct).  If 
co  =  b  dx  —  a  dy  +  xy  dz,  calculate  J  co. 

7.  Evaluate  fc  co,  where  C  is  the  unit  circle  x2  +  y2  =  1, 
oriented  counterclockwise,  and  co  =  y  dx  —  x  dy. 

8.  Compute  fc  co,  where  C  is  the  line  segment  in  R" 
from  (0,  0,  ... ,  0)  to  (3,  3,  ... ,  3)  and  co  =  x\  dx\  + 


x\  dx2  +  •  •  •  +  x"  dxn . 


9.  Evaluate  the  integral  f„  co,  where  X  is  the  parametrized 
helicoid 

X(s,  f)  =  (s  cosf,  s  sinr,  t),  0  <  s  <  1,  0  <  t  <  An 
and 

co  =  z  dx  A  dy  +  3  dz  A  dx  —  x  dy  A  dz. 

10.  Consider  the  helicoid  parametrized  as 

X(mi ,  ui)  =  (u\  cos3m2,"i  sin3M2,  5M2), 

0  <  ui  <  5,  0  <  m2  <  lit. 

Let  S  denote  the  underlying  surface  of  the  helicoid  and 
let  Q,  be  the  orientation  2-form  defined  in  terms  of  X 


X(Ml,«2 


)(a,  b)  =  det 


— 5sin3«2  a\  b\ 
5  cos  3;<2    a-2  b2 

—  3u\        fl3  £>3 


(a)  Explain  why  the  parametrization  X  is  incompatible 
with  Q. 

(b)  Modify  the  parametrization  X  to  one  having  the 
same  underlying  surface  5  but  that  is  compatible 
with  n. 

(c)  Alternatively,  modify  the  orientation  2-form  Q  to  £2' 
so  that  the  original  parametrization  X  is  compatible 
with  £2'. 


(d)  Calculate   fs  co,   where   co  =  z  dx  A  dy  —  (x1  + 
y2)dy  A  dz  and  S  is  oriented  using  £2. 

11.  Let  M  be  the  subset  of  R3  given  by  {{x,  y,  z)  \  x2  + 
y2  —  6  <  z  <  4  —  x2  —  y2}.  Then  M  may  be 
parametrized  as  a  3 -manifold  via 

X:  D— > R3;  X(i<i ,  u2, 113)  =  (u\  cos u2,u  \  sini<2,  M3), 

where 

D  =  {(uu  u2,  m3)  e  R3  ]  0  <  in  <  VI,  0  <  u2  <  lit, 

u\  -  6  <  m3  <  4  -  u\}. 

(The  parameters  u\,  u2,  and  M3  correspond,  respec- 
tively, to  the  cylindrical  coordinates  r,  9,  and  z.  Hence, 
it  is  straightforward  to  obtain  the  aforementioned 
parametrization. ) 

(a)  Orient  M  by  using  the  3-form  Q.,  where 

^x(u)(a,  b,  c)  =  det  [a  b  cl. 

Show  that  the  parametrization,  when  smooth,  is 
compatible  with  this  orientation. 

(b)  Identify  BM  and  parametrize  it  as  a  union  of  two 
2-manifolds  (i.e.,  as  a  piecewise  smooth  surface). 

( c)  Describe  the  outward-pointing  unit  vector  V,  vary- 
ing continuously  along  each  smooth  piece  of  3M, 
that  is  normal  to  BM.  Give  formulas  for  it  in  terms 
of  the  parametrizations  used  in  part  (b). 

1 2.  Calculate  fs  co,  where  S  is  the  portion  of  the  paraboloid 

z  =  x2  +  y2  with  0  <  z  <  4,  oriented  by  upward- 
pointing  normal  vector  (— 2x,  —2y,  1),  and  co  = 
ezdx  A  dy  +  y  dz  A  dx  +  x  dy  A  dz. 

13.  Calculate  fs  co,  where  S  is  the  portion  of  the  cylinder 
4  with  —  1  <  y  <  3,  oriented  by  outward 


xz  +  zz 


normal  vector  (x ,  0,  z),and&>  =  z  dx  A  dy  +  ey  dz  A 
dx  +  x  dy  A  dz. 

1 4.  Consider  the  parametrized  2-manifold 

X:  [1,  3]  x  [0,  2tt)  —>  R4,  X(s,t) 

=  (-y/s  cos  t,  V4  —  s  sin  t ,  -v/s  sin  t ,  -J A  —  s  cos  t). 

Find 


{x\  +  X4)  dx\  A  dx->,  —  (2x\  +  2x|)  dx2  A  dxs,. 


15.  Consider  the  parametrized  3-manifold 
X:  [0,  1]  x  [0,  1]  x  [0,  1]  R4, 


X(«l,  U2,  M3)  =  (Ml,  M2,  M3,  (2u\  —  M3)  ). 


Find 


L 


X2  dx2  A  JX3  A  dx4  +  2xiX3  Jxi  A  dx2  A  rfx3. 


8.3  I  The  Generalized  Stokes's  Theorem  553 


8.3  The  Generalized  Stokes's  Theorem 

We  conclude  with  a  discussion  of  a  generalization  of  Stokes's  theorem  that  relates 
the  integral  of  a  £-form  over  a  k -manifold  to  the  integral  of  a  (k  —  l)-form  over 
the  boundary  of  the  manifold.  Before  we  may  state  the  result,  however,  we  need 
to  introduce  the  notion  of  the  exterior  derivative  of  a  &-form. 

The  Exterior  Derivative  — 

The  exterior  derivative  is  an  operator,  denoted  d,  that  takes  differential  &-forms 
to  (k  +  l)-forms  and  is  defined  as  follows: 


DEFINITION  3.1    The  exterior  derivative  df  of  a  0-form  /  on  U  C  R"  is 
the  1-form 

At        df  A      j.^  A  V  A 

df  =  - —  dx\  +  - —  dx2-\  h  - —  dxn. 

ax\  0x2  oxn 

For  k  >  0,  the  exterior  derivative  of  a  &-form 

co  =      Fk_ikdxh  A  ■  •  •  A  dxh 

is  the  (k  +  l)-form 

dco  =  J^JjdFju,jk)  A  dx{x  A  ■  •  •  A  dxit, 

where  dFiujt  is  computed  as  the  exterior  derivative  of  a  0-form. 
EXAMPLE  1  If 

f(xl,x2,xi,x4,x5,x6)  =  X\X2Xi>  +  X4X5X6, 

then 

df  =  x2xt,  dx\  +        dx2  +  X\x2  dxj,  +  x^Xf,  dx4  +  X4X6  dx$  +  X4X5  cfxg.  ♦ 

EXAMPLE  2   If  &)  is  the  1-form 

00  =  x\x2  dx\  +  X2X3       +  (2xi  —  x2)  dxj,, 

then 

Jo)  =  d(x\x2)  A  Jxi  +  d(x2Xi)  A  Jx2  +  d(2xi  —  x2)  A  <ix3 

=  (x2  dx\  +  xi        A  dx\  +  (X3  J%2  +  X2  dx?,)  A  Jx2  +  (2dxi  —  dx2)  A  dx$ . 

Using  the  distributivity  property  in  Proposition  1 .4  and  the  facts  that  dxj  A  dx\  =  0 

and  dxj  A  rfx;-  =  —  dxj  A  dx,-,  we  have 

da>  =  x\  dx2  A  dx\  +  X2  dx3  A  dx2  +  2  Jxi  A  dx^  —  dx2  A  JX3 

=  —  x\  dx\  A  dx2  +  2  Jxi  A  JX3  —  (X2  +  1) dx2  A  dxj,.  ^ 

Stokes's  Theorem  for  k -forms   

We  now  can  state  a  generalization  of  Stokes's  theorem  to  smooth  parametrized 
k -manifolds  in  R" . 

THEOREM  3.2  (Generalized  Stokes's  theorem)  Let  D  c  Rk  be  a  closed 
bounded,  connected  region,  and  let  M  =  X(D)  be  an  oriented,  parametrized  k- 
manifold  in  R".  If  dM  ^  0,  let  3M  be  given  the  orientation  induced  from  that 


554       Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


of  M.  Let  a)  denote  a  (k  —  l)-form  denned  on  an  open  set  in  R"  that  contains  M. 
Then 


/  doo  =  I  i 

JM  JdM 


If  9M  =  0,  then  we  take  faM  co  to  be  0  in  the  preceding  equation. 

We  make  no  attempt  to  prove  Theorem  3.2. 1  Instead,  we  content  ourselves 
for  the  moment  by  checking  its  correctness  in  a  particular  instance. 

EXAMPLE  3   We  verify  the  generalized  Stokes's  theorem  (Theorem  3.2)  for  the 
2-form  co  =  zw  dx  A  dy,  where  M  is  the  3-manifold  M  =  {(x,  y,  z,  10)  e  R4 
w  =  x2  +  y2  +  z2,  x2  +  y2  +  z2  <  1}  oriented  by  the  3-form  £2  corresponding 


to  the  unit  normal 


N  = 


(2x,2y,2z,  -1) 
y^x2  +  4y2  +  4z2  +  1 


The  manifold  M  is  a  portion  of  the  3-manifold  given  in  Example  7  of  §8.2 
and  may  be  parametrized  as 


X:B 


R 


X(mj,  U2,  M3)  =  (U\,  U2,  «3,  u\  +  U2  +  uf), 


where  B  =  {(u\,  u%,  M3)  |  u\  +  u\  +  u\  <  1}.  Using  this  parametrization,  we 
have 

TM1  =  (1,0,0, 2u{) 
T„2  =  (0,  1,0,  2u2) 
T„3  =(0,0,  1,2k3) 

(2m  1,  2«2,  2«3,  —1) 


N  = 


4w2  +       +       +  1 


so  the  orientation  3-form  Q  is  given  by 

^x(u)(ai,  a2,  a3)  =  det  [N  ai   a2  a3]. 

Example  7  of  §8.2  shows  that  the  parametrization  X  is  compatible  with  this 
orientation.  Hence,  we  may  use  this  parametrization  without  any  adjustments 
when  we  calculate  fM  dco. 

The  boundary  of  M  is  dM  =  {(x,  y,  z,  w)  \  x2  +  y2  +  z2  =  w  =  1}  and  may 
be  parametrized  as 


Y:  [0,  tt]  x  [0,  2n) 
Then 

and 


T  — 

-1  Si  — 


R4,     Y(ji,  s2)  =  (sin^!  C0S52,  sins\  sms2,  cos^i,  1). 
(cos^i  cos^2.  cos  ifi  sin.?2,  —  sinsi,  0) 
(—  sinii  sin^2,  sin^i  cos^,  0,  0). 


For  a  full  and  rigorous  discussion  of  differential  forms  and  the  generalized  Stokes's  theorem,  see  J.  R. 
Munkres,  Analysis  on  Manifolds  (Addison- Wesley,  1991),  Chapters  6  and  7. 


8.3  I  The  Generalized  Stokes's  Theorem 


555 


An  outward-pointing  unit  vector  V  =  (vi,  v2,  v-$,  V4)  tangent  to  M  and  normal  to 
dM  must  satisfy 

•  V-N  =  0  along  dM; 

•  V-T.S1  =V-T.S2  =0. 

Along  dM,  we  have 

N  =  — =(2sinsi  coss2,  2sinsi  sins2,  2cossi,  —1). 
v  5 

Thus,  V  must  satisfy  the  system  of  equations 

(2sinsi  coss2)i>i  +  (2  sin  s\  sin  52)^2  +  (2cossi)i>3  —  V4  =  0 

(cossi  COSS2M  +  (cos si  sins2)u2  —  (sinsi)^  =  0 

— (sinsi  sins2)i>i  +  (sinsi  coss2)i>2  =  0 

After  some  manipulation,  one  finds  that  the  unit  vector  that  satisfies  these  equa- 
tions and  also  points  away  from  M  is 

V  =  — =(sinsi  cos S2,  sinsi  sins2,  cossi,  2). 
V  5 

Then  the  induced  orientation  2-form  Q3M  for  dM  is  given  by 

^)(ai,a2)  =  nX(„)(V)ai,a2)) 
where  X(u)  =  Y(s).  In  particular,  we  have 
^)(T.sl,TS2)  =  det[N  V  T„  Tj 


=  det 


r^sinsicoss2  ^sinsicoss2  COSS1COSS2  —  sinsisins2 
■^sinsisins2   4|  sins  1  sin s2  cossisins2  sinsiCOSs2 


v/5 


COS  Si 

_  J_ 


COS  Si 

V5 


-sin  si 
0 


0 
0 


=  sin  si  >  0 

for  0  <  si  <  Tt.  Hence,  the  parametrization  Y  of  dM,  when  smooth,  is  compatible 
with  the  induced  orientation,  so  we  may  use  this  parametrization  to  calculate 


a  m 


CO. 


Now  we  are  ready  to  integrate.  We  first  compute  fM  da>.  Since  co  =  zw  dx  A 
dy,  we  have 


dm  =  d{zw)  A  dx  A  dy  =  (z  dw  +  w  dz)  A  dx  Ady 
=  z  dw  A  dx  A  dy  +  w  dz  A  dx  A  dy. 


Thus, 


/  dco=  dwx(u)(TUl,  T„2,  Tm)dui  du2du3 

Jm         J  J  Jb 


-III 


M3  dw  A  dx  A  dy(Tm ,  TH2,  T„3) 


+  {u\  +  u\  +  u\)dz  A  dx  A  dy(TUl,  T„2,  T„3)}  du\  du2  duj, 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Iff, 


"3 


lit 
1 

0 


0 
1 

=2l(2 


+  (Mj  +  u\  +  u\) 


0  0  1 

1  0  0 
0      1  0 


=  1 


du\  du2  duj, 


J  J  J  (u\  + u\  +  3u\)du\dii2  duT,. 


Since  B  is  a  solid  unit  ball,  the  easiest  way  to  evaluate  this  iterated  integral  is  to 
use  spherical  coordinates  p,  (p,  and  9.  Hence, 


n  n7.7t      p7t  n\ 

I  dco=  I  I  I  (p2  +  2p2  cos2  cp)  p2  sirup  dpdcpdO 
Jm         Jo    Jo  Jo 

=  /      /     /    p4  (sirup  +  2  cos2  <p  simp)  dp  dtp  dd 
Jo    Jo  Jo 

f2n  r  1  /  7  \ 

=  J      /    -(sin<p+2cos  cp  sirup)  dcp  d9 

Jo    Jo  5 

-f 

5  Jo 

f 

Jo 


-lie 

1  f2n  10 


coscp  —  -  cos  <p 


4n 

dG  =  — . 
5  7n      3  3 


On  the  other  hand 


/    m  =  /  /  tBY(S)(Tft,  TSz)dsids2 

JSM  J  J[0,jr]x[0,2jr) 

p2tz  pTZ 

=  I      I    cos  S\  dx  A  dy(TSl ,  TS2)  ds\  ds2 
Jo  Jo 

p2tz  pit 

Jo  Jo 


COS  Si 


cos  S\  cos  52  —  sin  si  sin  s2 
cos  si  sin  S2      sin  si  cos  S2 


ds\  ds2 


PZJV  pit 

=         /    cossi  (cossi  sinsi)  ds\  d$2 
Jo  Jo 

=  /  / 

Jo  Jo 


cos  si  sin  si  Jsi  ds2 


p2n    /     y  \  n  /*2jr 


2  An 

-  ds2  =  —  • 

3  3 


Therefore,  the  generalized  Stokes's  theorem  is  verified  in  this  case. 


8.3  I  The  Generalized  Stokes's  Theorem  557 


Besides  being  notationally  elegant,  the  integral  formula  in  Theorem  3.2  beau- 
tifully encompasses  all  three  of  the  major  results  of  vector  analysis,  as  we  now 
show. 

First,  let  co  be  a  1-form  defined  on  an  open  set  U  in  R2.  Then 


co  =  M(x,  y)dx  +  N(x,  y)dy, 


so  that 


dco  =  dM  A  dx  +  dN  A  dy 


dM  J       dM     \     J       (dN  J       dN  J 

 dx  H  dy  I  A  dx  +  I   dx  H  dy  )  A  dy 

dx  dy      J  \  dx  dy 


dN  dM 


dx 


dy 


dx  A  dy. 


The  generalized  Stokes's  theorem  (Theorem  3.2)  says  that  if  D  is  a  2-manifold 
contained  in  U  and  dD  is  given  the  induced  orientation  (see  Figure  8.9),  then 


/  dco  =  I  co, 

J  D  JdD 


or,  in  this  instance,  that 


*  d  \  dx 
which  is  Green's  theorem. 


dN  dM\ 


dxdy=(t  Mdx  +  Ndy, 
dy  J  JdD 


= !L  (f  -  f ) dx  dy  L°> = LM  dx+Ndy 

Figure  8.9  The  generalized  Stokes's  theorem  implies  Green's  theorem. 

Next,  suppose  co  is  a  1-form  defined  on  an  open  set  U  in  R3.  Then 

co  =  F\(x,  y,  z)dx  +  F2(x,  y,z)dy  +  F3(x,  y,  z)dz- 
It  follows  that 

,'9*3     3F2\                /9Fj  3F3\ 
flftj  =  I  I  ay  A  dz  +  I  \  dz  Adx 

dy       dz  J  \  3z  3^ 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


Recall  from  Proposition  2.2  that  if  5  is  a  parametrized  2-manifold  (surface  in  R3), 
then 


/   co  =  (L  } 

JdS  JdS 


F-ds, 


where  F  =  F\  i  +  F2  j  +  F3  k.  From  Proposition  2.4, 

/  dco  =  I  I  G-dS, 
Js         J  Js 

where 


G=(^_^Wf---Vf---V=VxF. 

\  dy       dz  )       \  dz       dx  J       \  dx       by  J 

Theorem  3.2  tells  us,  if  S  is  oriented  and  35  is  given  the  induced  orientation,  that 


or,  equivalently,  that 


/   co  =  I  dco, 
Jas  Js 

(h  F-dS=  J  J  VxF-dS, 

JdS  J  J S 


which  is  the  classical  Stokes's  theorem.  (See  Figure  8.10.) 


J  dco  =  JJ  V  x  F  ■  dS 


Ls(o=LF-ds 


Figure  8.10  The  generalized  Stokes's  theorem  gives  the  classical  Stokes's 
theorem. 

Finally,  let  co  be  a  2-form  defined  on  an  open  set  in  R3 .  So 

co  =  F\(x,  y,  z)dy  A  dz  +  F2(x,  y,  z)dz  A  dx  +  F3(x,  y,  z)dx  A  dy. 
You  can  check  that 

SdFi     dF2  dF3\ 

dco  =  I  1  1  I  dx  A  dy  A  dz. 

\  dx       dy       dz  ) 

If  D  is  a  region  in  R3,  then  D  is  automatically  a  parametrized  3-manifold  since 
the  map  X:  D  —>  R3,  X(x,  y,  z)  =  (x,  y,  z)  parametrizes  D.  (One  can  show  that 
in  this  instance  D  is  always  orientable  as  well.)  If  D  is  bounded  and  3D  (which  is 
a  surface)  is  given  the  induced  orientation  (i.e.,  outward-pointing  normal),  then 


8.3  I  The  Generalized  Stokes's  Theorem  559 


Proposition  2.4  states  that 

f   co  =  £  F-dS, 

JdD  JJ3D 

where  F  =  F\  i  +  Fi  j  +  F3  k.  From  Example  6  of  §8.2, 

[  dco=  f  (—  +  — +  —  )dx  Ady  Adz=  SIS  V-FdV. 
Jd  JD\dx       dy       dz  J  J  J  JD 

Theorem  3.2  indicates  that  fgD  co  =  fD  dco  or 

(jf  F-dS  =  [  [  [  V-FdV, 

JJdD  J  J  Jd 

which  is,  of  course,  Gauss's  theorem.  (See  Figure  8.11.) 


Figure  8.1 1  The  generalized  Stokes's  theorem  gives 
rise  to  Gauss's  theorem. 


In  the  foregoing  remarks,  we  have  implicitly  set  up  a  sort  of  "dictionary" 
between  the  language  of  differential  forms  and  exterior  derivatives  and  that  of 
scalar  and  vector  fields.  To  be  explicit,  see  the  table  of  correspondences  shown  in 
Figure  8.12. 

The  theorems  of  Green,  Stokes,  and  Gauss  all  arise  from  Theorem  3.2 
applied  to  1 -forms  and  2-forms.  The  next  question  is,  can  the  "dictionary"  and 
Theorem  3.2  provide  a  corresponding  result  for  0-forms?  The  generalized  Stokes's 
theorem  (Theorem  3.2)  states,  for  a  0-form  co  and  an  oriented  parametrized  curve 
C,  that 

/  dco  =  I  co. 

Jc  Jbc 


k 

Differential  &-form 

Field 

Derivative 

0 

Scalar  field  / 

dco  -0-  V/ 

1 

co  =  F\  dx  +  F2  dy  +  Ft,  dz 

Vector  field 

dco  •*»  V  x  F 

F  =  Fx  i  +  F2  j  +  F3  k 

2 

co  =  F\  dy  A  dz  +  Fi  dz  A  dx  +  F3  dx  A  dy 

Vector  field 

dco  +±  V  •  F 

F  =  Fx  i  +  F2  j  +  F3  k 

Figure  8.1 2  A  differential  forms-vector  fields  dictionary. 


Chapter  8  i  Vector  Analysis  in  Higher  Dimensions 


Now,  if  C  is  closed,  then  3C  is  empty  (and  so  fac  co  =  0).  But  if  C  is  not  closed, 
then  3C  consists  of  just  two  points.  In  that  case,  what  should  fgc  a>  mean?  In 
particular,  to  apply  Theorem  3.2,  we  must  orient  dC  in  a  manner  that  is  consistent 
with  the  orientation  of  C,  which  can  be  done  by  assigning  a  "— "  sign  to  the  initial 
point  A  of  C  and  a  "+"  sign  to  the  terminal  point  B.  (See  Figure  8.13.)  Then 
fdC  co  is  just  f(B)  —  f(A),  where  /  is  the  function  (scalar  field)  corresponding 
to  co  in  the  table.  Since  dco  corresponds  to  V /,  Theorem  3.2  tells  us  that 

f  Vf.di  =  f(B)-f(A),  (1) 

the  result  of  Theorem  3.3  in  Chapter  6. 

Finally,  for  the  case  n  =  1,  that  is,  the  case  of  0-forms  (functions)  on  R,  the 
0-form  co  corresponds  to  a  function  /  of  a  single  variable,  and  V/  is  the  ordinary 
derivative  /'.  Furthermore,  a  parametrized  curve  in  R  is  simply  a  closed  interval 
[a,  b].  Then  equation  (1)  reduces  to 

b 

f(x)dx  =  f(b)-f(a), 

a  version  of  the  fundamental  theorem  of  calculus.  Thus,  we  can  appreciate  that 
the  generalized  Stokes's  theorem  is  an  elegant  and  powerful  generalization  of  the 
fundamental  theorem  of  calculus  to  arbitrary  dimensions. 

8.3  Exercises 


Figure  8.13  The  orientation 
of  the  curve  C  induces  an 
orientation  of  its  boundary 
(i.e.,  the  endpoints  A  and  B). 


In  Exercises  1-7,  determine  dco,  where  co  is  as  indicated. 


1. 

CO  = 

exyz 

2. 

CO  = 

x3y  —  2xz2  +  xy2z 

3. 

CO  = 

(x2  +  y2)  dx  +  xy  dy 

4. 

CO  = 

x\  dx2  —  X2  dx\  +  X1X4  dx4  —  X4X5  dxs 

5. 

CO  = 

xz  dx  A  dy  —  y2z  dx  A  dz 

6. 

CO  = 

X1X2X3  dx2  A  dxi  A  dx4  +  X2JC3X4  dx\  A  dx2 

A 

dxT, 

7. 

CO  = 

Yll=i  xf  dx\  A  •  •  •  A  dxj  A  •  •  •  A  dxn  (Note: 

dxi 

means  that  the  term  dxt  is  omitted.) 


8.  Let  u  be  a  unit  vector  and  /  a  differentiable  function. 
Show  that  dfXo(u)  =  Duf  (x0).  (Recall  that  Du/(x0) 
denotes  the  directional  derivative  of  /  at  xo  in  the  di- 
rection of  u.) 

9.  If  co  =  F(x,  z)  dy  +  G(x,  y)  dz  is  a  (differentiable)  1- 
form  on  R3,  what  can  F  and  G  be  so  that  dco  = 

z  dx  A  dy  +  y  dx  A  dz7 

1 0.  Verify  the  generalized  Stokes's  theorem  (Theorem  3.2) 
for  the  3-manifold  M  of  Exercise  11  of  §8.2,  where 
&)  =  2x  dy  A  dz  —  Z  dx  A  dy. 


1 1 .  Verify  the  generalized  Stokes's  theorem  (Theorem  3.2) 
for  the  3-manifold 

M={(x,  y,  z,  ui)eR4|  x=S  -  2y2  -  2z2  -  2w2,  x  >  0} 

and  the  2-form  co  =  xy  dz  A  dw.  (Hint:  First  compute 
f3M  co.  To  calculate  fM  dco,  study  Example  3  of  this 
section.) 

12.  (a)  Let  M  be  a  parametrized  3-manifold  in  R3  (i.e.,  a 

solid).  Show  that 

1  f 

Volume  of  M  =  —  I    x  dy  A  dz  —  y  dx  A  dz 
3  Jsm 

+  z  dx  A  dy. 

(b)  Let  M  be  a  parametrized  n  -manifold  in  R".  Ex- 
plain why  we  should  have 

n -dimensional  volume  of  M 

1  f 

=  —  I    X\  dx2  A  •  •  •  A  dx„ 
n  Jsm 

—  X2  dx\  A  dxi  A  •  •  ■  A  dxn 

+  Xi  dx\  A  dx2  A  dX4  A  •  •  •  A  dx„  +  •  •  • 

+  (—  \)"~  xn  dx\  A  dx2  A  •  •  •  A  dxn-\. 


Miscellaneous  Exercises  for  Chapter  8 


True/False  Exercises  for  Chapter  8 


1.  (dx  Ady  +  dy  Adz)((l,0,  1),  (0,-1,3))=  0. 

2.  dx\  A  d\2  A  dxi  A  dx4  =  dx2  A  dx$  A  dx\  A  c/.^. 

3.  There  are  21  basic  5-forms  in  R7. 

4.  rfxi  A  dx2  =  dx2  A  dx\. 

5.  (t/jti  A  </X2)  A  dx?,  =  dxi  A  (dx\  A  c/-t2)- 

6.  If  a)  is  a  3-form  on  R6  and  r]  is  a  5-form  on  R6,  then 

d)M)  =  l)AO). 

7.  If  a)  is  a  2-form  on  R8  and  )?  is  a  3-form  on  R8,  then 

CO  A  Tj  =  T]  A  CO. 

8.  dx  Ady  A  dz(a,  b,  c)  =  — c/z  A  dy  A  dx(a,  c,  b). 

9.  dxj  a  dxj(a,  b)  =  —dx,  a  dxj(b,  a). 

10.  Let  D  =  [0, 2]  x  [-1,  1]  and  let  X:  D      R4  be  given 
by 

X0,0  =  0-f,  Jf2,  se',4t). 

Then  M  =  X(D)  is  a  smooth  parametrized  2-manifold 
inR4. 

11.  Let  D  =  [-2,  2]  x  [0,  5]  x  [-3,  3]  and  let  X:  D 
R4  be  given  by 

X(«i,  M2,  M3)  =  (u\ll\,  «2COSM3'  U\  —  U2,  u\u\). 

Then  M  =  X(£>)  is  a  smooth  parametrized  3 -manifold 
inR4. 

12.  If  D  =  [0,  1]  x  [0,  1],  then  the  underlying  manifolds 
of  X:  D  R3, 

X(s,  t)  =  (s  coslirt,  .ysin2jr?,  s2) 
andY:D  ->  R3, 

Y(s,  t)  =  (t  cos27rj,  fsin27rs,  f2) 
are  the  same. 

13.  Let  co  =  dxAdy   and   D  =  [0,  1]  x  [0,  1].  Then 
J  a)  =  jY  co,  where  X:  D  — >  R3, 

X(s,  t)  =  (s  coslirt,  ssin2jtt,  s2), 


andY:D  -*  R3, 

Y(s,  t)  =  (tcoslirs,  ?sin27rs,  f2). 

14.  Let  5  =  {u  e  R3  |  u\  +  u\  +  u\  <  1}.  The  general- 
ized paraboloid  X:  B      R4  defined  by 

X(u\,  U2,  M3)  =  (u\,  142,  M3,  M j  +  2l<2  +  3k2) 

has  as  its  boundary  the  ellipsoid  Y:  [0,  jt]  x  [0,  27r)  — > 

Y(s,  t)  =  (sins  cost,  A=  sins  sin t,  -Lcoss,  1). 

15.  Let  M  C  R"  be  the  graph  of  a  function  f:UQ 
R""1  -»■  R"  parametrized  by  X:  U  ->  R", 

X(Mi,  .  .  .  ,  M„_i)  =  («!,...,  M„_i,  /(Mi,  .  .  .  ,  M„_i)). 
If 

TVT/  \  (/"I  fun-1  '  ~  1) 

N(«l,  .  .  .  ,  = 

y/;21+---  +  /IL  +  1 

is  a  unit  normal,  then  the  parametrization  X  is  compat- 
ible with  the  (n  —  1  )-form  f2  defined  by 

^x(U)(ai , . . . ,  a„_i)  =  det  [  ai   •  •  •   a„_i   N  ] . 

16.  If  co  =  x-[Xi  dx2  A  dx4,  then  dco  =  X3  dx\  A  dx2  A 
dx4  +  x\  dx2  A  dxj  A  dx4. 

17.  If  co  =  x\  dxi  —  X2  dx\  +  x\X2Xt,  dxi,  then 

dco  =  (X2X3  +  \)dx\  A  dxi  +  dx\  A  dx2 
+  X1JC3  dx2  A  dx?,. 

18.  If  co  =  x\X2  dx\  A  dx2  +  X2X3  dx\  A  dx-$  +  x^x?  dx2 
Adxi,  then 

dco  =  2x2  dx\  A  dx2  A  dx$. 

19.  If  co  is  an  H-form  on  R",  then  dco  =  0. 

20.  If  M  is  a  parametrized  fc-manifold  without  boundary 
in  R"  and  co  is  (k  —  1  )-form  defined  on  an  open  set 
containing  M,  then  fM  dco  =  0. 


Miscellaneous  Exercises  for  Chapter  8 


1.  Let  co  be  a  £-form,    an  /-form.  Show  that 

d(co  A  rj)  =  dco  At]  +  (—  l)kco  A  dr\. 

This  is  accomplished  by  the  following  steps: 
(a)  Show  that  the  result  is  true  when  k  =  I  =  0,  that  is, 
when  co  =  f  and  ii  =  g.  (Here  /  and  g  are  scalar- 
valued  functions.) 


(b)  Establish  the  result  when  k  =  0  and  /  >  0. 

(c)  Establish  the  result  when  k  >  0  and  /  =  0. 

(d)  Establish  the  result  when  k  and  /  are  both  positive. 

2.  Let    M    be    the    subset    of  R5    described  as 

{(x\,  X2,  Xt,,  X4,  X5)  |  X5  =  X1X2X3X4,      0  <  X\,  X2,  Xl, 

x4  <  1}. 


Chapter  8  |  Vector  Analysis  in  Higher  Dimensions 


(a)  Give  a  parametrization  for  M  (as  a  4-manifold) 
and  check  that  your  parametrization  is  compati- 
ble with  the  orientation  4-form  f2  =  dx\  A  dx2  A 
dxi  A  dx$. 

(b)  Calculate  JM  X4  dx\  A  <ijC2  A  dx^  A  J.«5. 

3.  (a)  Let  C  be  the  curve  in  R2  given  by  y  =  f(x), 
a  <  x  <b.  Assume  that  /  is  of  class  C1.  If  C 
is  oriented  by  the  direction  in  which  x  increases, 
show  that  if  co  =  y  dx,  then 


L 


area  under  the  graph  of  /. 


(b)  Let  S  be  the  surface  in  R3  given  by  the  equation  z  = 
f(x,  y),  where  (x,  y)  e  [a,  b]  x  [c,  d].  Assume 
that  /  is  of  class  C  .  If  S  is  oriented  by  upward- 
pointing  normal,  show  that  if  co  =  z  dx  A  dy,  then 


volume  under  the  graph  of  / . 


(c)  Now  we  generalize  parts  (a)  and  (b)  as  follows: 
Suppose  f-.D—*  R  is  a  function  of  class  C1  de- 
fined on  a  connected  region  D  c  R"_1 .  Let  M  be 
the  (n  —  l)-dimensional  hypersurface  in  R"  de- 
fined by  the  equation  x„  =  f(x\,  . . . ,  *„-i),  where 
(jci,  . . . ,  G  D.  If  co  =  x„  dx\  A  •  •  •  A  dxn-\, 
show  that 


JM 


co  =  ±(«-dimensional  volume 


under  the  graph  of  /). 
How  can  we  guarantee  a  "+"  sign  in  the  equation? 

4.  Let  M  be  the  portion  of  the  cylinder  x2  +  z2  =  1, 
0  <  y  <  3,  oriented  by  unit  normal  N  =  (x,  0,  z). 

(a)  Use  N  to  give  an  orientation  2-form  £2  for  M.  Find 
a  parametrization  for  M  compatible  with  Q. 

(b)  Identify  9M  and  parametrize  it. 

(c)  Determine  the  orientation  form  Q3M  for  3M  in- 
duced from  n  of  part  (a). 

(d)  Verify  the  generalized  Stokes's  theorem  (Theorem 
3.2)  for  M  and  co  =  zdx  +  (x  +  y  +  z)dy—x  dz. 

5.  Use  the  generalized  Stokes's  theorem  to  calcu- 
late fs4  co,  where  S4  denotes  the  unit  4-sphere 
l(xi ,  x2,  xi,X4,  X5)  G  R5  I  x\  +  x\  +  x\  +  x\  +  x\  = 
1}  and  co  =  x^  dx\  A  dx2  A  dx4  A  dxs  +  X4  dx\  A 
dx2  A  dxz  A  dxs . 

6.  (a)  Let  co  be  a  0-form  (i.e.,  a  function)  of  class  C2. 

Show  that  d(dco)  =  0. 

(b)  Now  suppose  that    is  a  &-form  of  class  C2,  mean- 
ing that  when  co  is  written  as 


Pi\  ...it  dxix  A       A  dXjk , 


each  Fi,.„i,  is  of  class  C2.  Use  part  (a)  and  the  result 
of  Exercise  1  to  show  that  d(dco)  =  0. 

7.  In  this  problem,  show  that  the  equation  d(dco)  =  0  im- 
plies two  well-known  results  about  scalar  and  vector 
fields. 

(a)  First,  let  co  be  a  0-form  (of  class  C2 ).  Then  co  corre- 
sponds to  a  scalar  field  / .  Use  the  chart  on  page  559 
to  interpret  the  equation  d(dco)  =  0. 

(b)  Next,  suppose  that  co  is  a  1-form  (again  of  class 
C2).  Then  co  corresponds  to  a  vector  field.  Inter- 
pret the  equation  d(dco)  =  0  in  this  case. 


8.  Let 


x  dy  A  dz  +  y  dz  A  dx  +  z  dx  A  dy 


(x2  +  y2  +  z2fl2 

(a)  Evaluate  fs  co,  where  S  is  the  unit  sphere  x2  + 
y2  +  z2  =  1,  oriented  by  outward  normal. 

(b)  Calculate  dco. 


Figure  8.14  Figure  for  Exercise  ! 


(c)  Verify  Theorem  3.2  over  the  region  M  = 
{(x,  y,  z)  \  a2  <  x2  +  y2  +  z2  <  1},  where  a  /  0. 

(d)  Now  let  M  be  the  solid  unit  ball  x2  +  y2  +  z2  <  1 . 
Does  Theorem  3.2  hold  for  M  and  col  Why  or  why 
not? 

(e)  Suppose  that  S  is  any  closed,  bounded  surface  that 
lies  entirely  outside  the  sphere  S(  =  {(x,  y,  z)  \ 
x2  +  y2+z2  =  e2}.  (See  Figure  8.14.)  Argue  that  if 
S  is  oriented  by  outward  normal,  then  fs  co  =  An . 

9.  Let  M  be  an  oriented  (k  +  l  +  1  )-manifold  in  R";  let 
&)  be  a  fc-form  and  r\  an  /-form  defined  on  an  open  set 
of  R"  that  contains  M.  If  dM  =  0,  use  Theorem  3.2 
and  Exercise  1  to  show  that 


/  dco  A  11 

JM 


(-1) 


Ml 


/  coAdrj. 

J  M 


10.  Let  M  be  an  oriented  k -manifold.  Use  Exercise  1  and 
the  general  version  of  Stokes's  theorem  to  establish 
"integration  by  parts"  for  &-forms  co  and  0-forms  /: 


l<ii  <---<it  <n 


f  fdco 

J  M 


/ 


fto 


I  df  A  co. 

JM 


Suggestions  for  Further 
Reading 


General 

Francis  J.  Flanigan  and  Jerry  L.  Kazdan,  Calculus  Two:  Linear 
and  Nonlinear  Functions,  2nd  ed.,  Springer,  1990.  The  essen- 
tials of  vector  calculus  are  presented  from  a  linear-algebraic 
perspective. 

John  H.  Hubbard  and  Barbara  Burke  Hubbard,  Vector  Calculus, 
Linear  Algebra,  and  Differential  Forms:  A  Unified  Approach, 
2nd  ed.,  Prentice  Hall,  2002.  Treats  the  main  topics  of  multi- 
variable  calculus,  plus  a  significant  amount  of  linear  algebra. 
More  sophisticated  in  approach  than  the  current  book,  using 
differential  forms  to  treat  integration. 

Jerrold  E.  Marsden  and  Anthony  J.  Tromba,  Vector  Calculus, 
5th  ed.,  W.  H.  Freeman,  2003.  This  text  is  probably  the  one  most 
similar  to  the  current  book  in  both  coverage  and  approach, 
using  matrices  and  vectors  to  treat  multivariable  calculus 
inR". 

Jerrold  E.  Marsden,  Anthony  J.  Tromba,  and  Alan  Weinstein, 
Basic  Multivariable  Calculus,  Springer/W.  H.  Freeman,  1993. 
Somewhat  similar  to  Marsden  and  Tromba's  Vector  Calculus, 
but  with  less  emphasis  on  a  rigorous  development  of  the  sub- 
ject. A  good  guide  to  the  main  ideas. 

George  B.  Thomas  and  Ross  L.  Finney,  Calculus  and  Ana- 
lytic Geometry,  9th  ed.,  Addison-Wesley,  1996.  A  complete 
treatment  of  the  techniques  of  both  single-variable  and  multi- 
variable  calculus.  No  linear  algebra  needed.  A  good  reference 
for  the  main  methods  of  vector  calculus  of  functions  of  two 
and  three  variables. 

Richard  E.  Williamson,  Richard  H.  Crowell,  and  Hale  F. 
Trotter,  Calculus  of  Vector  Functions,  3rd  ed.,  Prentice  Hall, 
1 972 .  A  smooth  and  careful  treatment  of  multivariable  calculus 
and  vector  analysis,  using  linear  algebra. 

More  Advanced  Treatments 

The  following  texts  all  offer  relatively  rigorous  and  theoretical 
developments  of  the  main  results  of  multivariable  calculus.  As 
such,  they  are  especially  useful  for  studying  the  foundations  of 
the  subject. 

Tom  M.  Apostol,  Mathematical  Analysis:  A  Modern  Approach 
to  Advanced  Calculus,  1st  ed.,  Addison-Wesley,  1957.  Look  at 
the  first  edition,  since  the  second  edition  contains  much  less 
regarding  multivariable  topics. 


R.  Creighton  Buck,  Advanced  Calculus,  3rd  ed.,  McGraw-Hill, 
1978.  Treats  foundational  issues  in  both  single-variable  and 
multivariable  calculus.  Uses  the  notation  of  differential  forms 
for  considering  Green's,  Stokes's,  and  Gauss's  theorems. 

Richard  Courant  and  Fritz  John,  Introduction  to  Calculus  and 
Analysis,  Vol.  Two,  Wiley-Interscience,  1974.  A  famous  and 
encyclopedic  work  on  the  analysis  of  functions  of  more  than 
one  variable  (Volume  One  treats  functions  of  a  single  variable), 
with  fascinating  examples. 

Wilfred  Kaplan,  Advanced  Calculus,  3rd  ed.,  Addison-Wesley, 
1984.  A  full  treatment  of  advanced  calculus  of  functions  of  two 
and  three  variables,  plus  material  on  calculus  of  functions  of  n 
variables  (including  some  discussion  of  tensors).  In  addition, 
there  are  chapters  on  infinite  series,  differential  equations,  and 
functions  of  a  complex  variable. 

O.  D.  Kellogg,  Foundations  of  Potential  Theory,  originally  pub- 
lished by  Springer,  1929.  Reprinted  by  Dover  Publications, 
1954.  A  classic  work  that  ventures  well  beyond  the  subject  of 
the  current  book.  The  writing  style  may  seem  somewhat  old- 
fashioned,  but  Kellogg  includes  details  of  certain  arguments 
that  are  difficult  to  find  anywhere  else. 

James  R.  Munkres,  Analysis  on  Manifolds,  Addison-Wesley, 
1991.  A  superbly  well  written  and  sophisticated  treatment  of 
calculus  in  R" .  Requires  a  knowledge  of  linear  algebra.  In- 
cludes a  full  development  of  differential  forms  and  exterior  al- 
gebra to  treat  integration.  For  advanced  mathematics  students. 

David  V  Widder,  Advanced  Calculus,  2nd  ed.,  originally  pub- 
lished by  Prentice-Hall,  1961.  Reprinted  by  Dover  Publica- 
tions, 1989.  Careful  treatment  of  differentiation  and  integra- 
tion of  functions  of  one  and  several  variables.  Chapters  on 
differential  geometry,  too. 

Physics  Oriented  Texts 

Mary  L.  Boas,  Mathematical  Methods  in  the  Physical  Sciences, 
2nd  ed.,  Wiley,  1983.  A  wide  variety  of  topics  that  extend  well 
beyond  vector  calculus.  For  the  student  interested  in  physics  as 
well  as  mathematics. 

Harry  F.  Davis  and  Arthur  D.  Snider,  Introduction  to  Vector 
Analysis,  6th  ed.,  William  C.  Brown,  1991.  A  detailed  and  rel- 
atively sophisticated  treatment  of  multivariable  topics.  Includes 
appendices  on  classical  mechanics  and  electromagnetism. 


563 


564       Suggestions  for  Further  Reading 


Edward  M.  Purcell,  Electricity  and  Magnetism,  2nd  ed., 
McGraw-Hill,  1985.  This  is  a  physics,  not  a  mathematics,  text. 
Provides  excellent  intuition  regarding  the  meaning  of  line  and 
surface  integrals,  differential  operations  on  vector  and  scalar 
fields,  and  the  significance  of  vector  analysis. 

Other 

Alfred  Gray,  Modern  Differential  Geometry  of  Curves  and  Sur- 
faces with  Mathematica® ,  2nd  ed.,  CRC  Press,  1998.  Not 
a  vector  calculus  text  by  any  means,  but  rather  a  delightful 


library  of  geometric  objects  and  how  to  understand  them  via 
Mathematica.  Some  of  the  differential  geometric  topics  are 
somewhat  remote  from  the  subject  of  the  current  book,  but  the 
many  illustrations  are  worth  viewing.  An  outstanding  aid  for 
developing  one's  visualization  skills. 

H.  M.  Schey,  Div,  Grad,  Curl,  and  All  That:  An  Informal  Text 
on  Vector  Calculus,  3rd  ed.,  W.  W.  Norton,  1997.  The  classic 
"alternative"  book  on  vector  analysis,  aimed  at  students  of  elec- 
tricity and  magnetism.  Brief,  but  well  done,  intuitive  account 
of  vector  analysis. 


Answers  to  Selected 
Exercises 


Chapter  1 
Section  1.1 

1.  (a)  y 


.(2,1) 


(b) 


(3,3) 


(c) 


(-1,?) 


3.  (a)  (2,  8) 

(b)  (-16,-24) 

(c)  (5,  15) 

(d)  (-19,25) 

(e)  (8,  -26) 


b  (translated) 


7.  (a)  AB  =  (-4,3, 


(b)  AC  =  (1,  1,3), 
(-4,3,-1) 

(c)  In  general,  we  have 


1),BA  =  (4,  -3,  1) 
SC  =  (5,  -2,4), 


AC  +  CB 


9.  x  =  14,  v  =  16,  z  =  8 
1 1 .  Because  b  =  5a. 

13.  (1,  2,  3,  4)  +  (5,  -1,2,  0)  =  (6,  1,  5,  4) 

2(7,6, -3,1)  =  (14,  12, -6,  2) 

a  +  b  =  (fli  +  b\ , . . . ,  an  +  bn);  kn  =  (ka\ , . . . ,  kan) 
15.  =  (1,-4, -1,1) 

17.  If  your  displacement  vector  is  a  and  your  friend's  is  b,  then 
b  —  a  is  the  displacement  vector  from  you  to  your  friend. 
21.  (b)  The  position  vectors  of  points  in  the  parallelogram 

determined  by  (2,  2,  1)  and  (0,  3,  2) 
23.  (a)  V5  units  per  minute 

(b)  (0,  -4) 

(c)  7  minutes 

(d)  No 

25.  (a)  (5,5,4) 

(b)  (-5,  -5,  -4) 

Fi  =  (4,  -4,  2)  and  F2  =  (-4,  4,  8) 


27 


Section  1.2 

1.  2i  +  4j 
5.  2i  +  4j 


3.  3i  +  7T j  -  7k 

7.  (9, -2,V2) 


565 


566       Answers  to  Selected  Exercises 


9.  (JT.-1) 

11.  (a)  b  =  2ai  +  a2 

(b)  b  =  — lai  +4a2 

(c)  Take  a  =  (bi  +  b2)/2,  c2  =  (&,  -  b2)/2. 
13.  a:  =  f  +  2,  j  =  3f  -  1,  z  =  5  -  6f 

15.  x  =  t  +  2,  y  =  -It  -  1 

17.  x  =  t  +  1,  j  =  4,  z  =  5  -  6f 

19.  ;ti  =  1  -  2t,  x2  =  5t  +  2,  x3  =  3f ,  x4  =  7t  +  4 

21 .  (a)  x=2t-l,y  =  7-t,z  =  5t  +  3 

(b)  x  =  5  -  5t,  y  =  At  -  3,  z  =  5t  +  4 

(c)  One  alternative  for  (a):  x  =  —2t  —  l,y  =  t  +  7, 
z  =  3  —  5t ;  one  alternative  for  (b):  x  =  5t,  y  =  1  — 
At,  z  =  9  -  5t 

(d)  For  (a):  (x  +  l)/2  =  (y  -  7)/(-l)  =  (z  -  3)/5; 

for  (b):  (x  -  5)/(-5)  =  (y  +  3)/4  =  (z  -  4)/5; 

for  (c):  (*  +  l)/(-2)  =  (y-7)/l  =  (z-3)/(-5) 
and.x/5  =  (y  -  l)/(-4)  =  (z  -  9)/(-5). 

23.  (*  -  7)/l  =  (y  +  9)/3  =  (z  -  6)/(-8) 

25.  x  =  3?  -  5,  y  =  7f  +  1,  z  =  -2t  -  10 

27.  Multiply  the  first  symmetric  form  by  —  \,  Then  add  —  \ 
to  each  "side." 

29.  No.  Setting  t  =  0  in  the  equations  for  l\  gives  the  point 
(2,  —7,  1).  To  have  x  =  2  in  l2,  we  must  take  t  =\.  But 
?  =  j  gives  the  point  (2,  —7,  2)  on  l2,  not  (2,  —7,  1). 

31 .  No,  they  do  not.  Note  that  x  >  —  1,  y  >  3,  and  z  <  1 
only.  The  parametric  equations  define  a  ray  with  endpoint 
(-1,3,1). 

33.  (-14,5,-18) 

35.  With  x  =  0:  (0,  f ,  \);  with  y  =  0:  (-f ,  0,  with 

z  =  0:  (7,17,0). 
37.  No  43.  No 

45.  (a)  Circle  a2  +  y2  =  4,  traced  once  counterclockwise.  If 
0  <  t  <  2n,  circle  is  traced  three  times. 

(b)  Circle  x2  +  y2  =  25,  traced  once  counterclockwise. 

(c)  Circle  x2  +  y2  =  25,  traced  once  clockwise. 

(d)  Ellipse  *2/25  +  y2/9  =  1,  traced  once  counterclock- 
wise. 

47.  x  =  (a  +  a6)  cos  9,  y  =  (a  +  a6)  sinS 

Section  1.3 

1 .  a  •  b  =  13,  ||a||  =  ^26;  ||b||  =  7T3 

3.  a  •  b  =  -44,  ||a||  =  V50;  ||b||  =  2^14 

5.  a-b  =  2,  ||a||  =  V26;  ||b||  =  ^3 

7.  2;r/3       9.  cos"1  (V2/V3")       11.  n/2 

13.  |(i  +  j)  15.  2k 

17.  (2i-j  +  k)/V6 

19.  V3(i  +  j-k) 

21 .  Yes,  if  a  •  b  =  0  or  a  =  b. 


25.  (a)  Work  =  (component  of  F)||Pgj| 

=  ||F||  cos6»|pe|  =  F-PQ 

(b)  2 
27.  10,000  ft-lb 

29.  cosa  =  3/5,  cos/3  =  0,  cosy  =  4/5 

33.  Hint:  The  diagonals  are  given  by  di  =  a  +  b  and  d2  = 

b  —  a,  where  a  and  b  determine  the  adjacent  sides  of  the 

parallelogram. 

Section  1.4 

1.  2  3.  -5 

5.  (31,-5,8)  7.  5k 

9.  -6i+14j  +  4k  11.  5730 

13.  c  is  parallel  to  the  plane  determined  by  a  and  b  (or  else  a 
is  parallel  to  b). 


15.  V 1002/2 
21.  (axb)-c 


17 

3^ 

Z34/2 

a\ 

0.2 

a3 

bi 

b2 

b3 

C\ 

C2 

C'3 

bi 

b2 

b3 

C\ 

C2 

C'3 

a\ 

0.2 

«3 

Expand  determinants  to  see  they  are  equal. 
23.  (b)  1 
25.  (a)  a  x  b 

(b)  2a  x  b/||a  x  b|| 

(e)  a  x  (b  x  c) 

(f)  (a  x  b)  x  c 
35.  (a)  20V3  ft-lb 

(b)  30V3  ft-lb 
37.  75V3  ft-lb  39.  400;r/3  in/min 

41 .  Rotation  about  an  axis  that  passes  through  the  rigid  body 

is  very  different  from  rotation  about  a  parallel  axis  that 

does  not  pass  through  the  body. 

Section  1.5 

1 .  x  -  y  +  2z  =  8 

3.  5x-4y  +  3z  =  25 

5.  5x  -  Ay  +  z  =  12 

7.  x-y  +  lz  =  5 

11.  3x  -  2y  -  z  =  3 


Answers  to  Selected  Exercises  567 


13.  x  =  13f 
15.  -4/3 
17.  x=2s  +  t  - 
1 9.  x  =  2s  +  5t 
21.  x  =  3s  +  3t 


2i>y 


-I4t+24,z 


-5t 


l,y  =  2  -  3s,  z  =  s  -  St  +  7 

-  1,  y  =  lOt  -  3s  +  3,  z  =  4s  +  It  -  2 

-  5,  y  =  10  -  3s  -  6t,  z  =  2s-2t  +  9 


23.  I9x  -  I6y  +  lz 
25.  31/V34 


59 


27.  25/V64T 
(b)  The  lines  must  intersect. 


5z  =  2±3V35 


29.  (a)  0 
31.  7T4/2 
35.  x  +  3y 

37.  Hint:  Consider  Example  8  in  §1.5  with  A  as  Pi  and  B 
as  P2. 

Section  1.6 


1. 


e„ 


1) 


1) 


(a)  ei  +  2e2  +  3e3  H  1-  ne„ 

(b)  ei  -  e3  +  e4  -  e6  H  h  e„_2  ■ 

(1,-2,  3,-4,  ...,(-l)n+1n) 

(a)  (3,-1,  ll,-l,...,2«  +  (-l)"+12n 

(b)  (-1,  7,  -1, . . . ,  2n  +  (-l)n2n  -  1) 

(c)  (-3,-9,-15,  ...,-6n  +  3) 

(d)  %/l  +  9  +  25  +  ---  +  (2n-  l)2 

(e)  2  -  12  +  30  +  •  •  •  +  (-l)"+12«(2n  - 

Hint:  Use  the  triangle  inequality. 

Hint:  Square  both  sides  of  the  equation  and  write  as  dot 
products. 

1 3.  Hyperplane  in  R5  passing  through  (1,  —2,  0,  4,  —1)  and 
perpendicular  to  the  vector  n  =  (2,  3,  —7,  1,  —5) 

15.  (a)  Total  cost  = 

(200,  250,  300,  375,  450,  500)  •  (xux2,  x3,x4,  x5,  x6) 
(b)  {x  e  R6  |  (200, . . . ,  500)  •  (xu  . . . ,  x6) <  100,000}. 
Budget  hyperplane  is  200.x  i  +  •  •  •  +  500*6  =  100,000. 


17. 


9 
11 


19. 


21.  -42 


5 

8 

-2 

5 

-4 

0 

15 

-9 

5 

0 

23.  -240 


25.  (a)  A  lower  triangular  matrix  is  an  n  x  n  matrix  whose 
entries  above  the  main  diagonal  are  zero, 
(b)  For  an  upper  triangular  matrix,  use  cofactor  expansion 
about  the  first  column  or  last  row. 

27.  -289 


31. 


1/2 
0 
0 


-1  1/2 
1  0 
0  -1 


37. 


14 


14 


J_ 
14 


39.  (32,46,  19,-35) 

Section  1.7 

1.  (1,1) 

5.  (2^2,  3jt/4) 

9.  (-1/2,  V3/2,-2) 

13.  (0,0,-2) 

17.  (2V2,  jt/6,  7jt/4) 

19.  (a) 


3.  (3,0) 

7.  (2  cos  2,  2  sin  2,  2) 
11.  (o,3V3/2,3/2) 
15.  (2,2tt/3,  13) 


d  =  nl2 


(b) 


21. 


23.  Cartesian:  v  =  2;  cylindrical:  r  sin  6  =  2;  surface  is  a  ver- 
tical plane. 

25.  Cartesian:  x  =  y  =  0;  spherical:  <p  =  0  or  jt;  object  is  the 
z-axis. 


35.  Note:  The  desired  region  is  that  between  the  two  spheres 
shown. 

z 


y 


37.  (a)  r  =  —f(0)  is  the  reflection  through  the  origin  of 
r  =  f(0). 

(b)  p  =  —  f{<p,  9)  is  the  reflection  through  the  origin  of 
p  =  f(cp,9). 

(c)  r  =  3f(9)  is  a  threefold  magnification  of  r  =  f{6). 

(d)  p  =  3f(<p,6)    is    a   threefold   magnification  of 
p  =  f(cp,9). 

41 .  i  =  sin<p  cos#  ep  +  cos  <p  cos  9  ev  —  sin  9  eg 
j  =  sin  <p  sin  9  ep  +  cos  cp  sin  9  ev  +  cos  9  eg 
k  =  cos  (pep  —  sin  0 

45.  (a)  Hint:  use  part  (a)  of  the  previous  exercise. 

47.  Hint:  use  part  (a)  of  the  previous  exercise. 


True/False  Exercises  for  Chapter  1 

1 .  False.  (The  corresponding  components  must  be  equal.) 
3.  False.  ((—4,  —3,  —3)  is  the  displacement  vector  from  P2 
to  Pi.) 

5.  False.  (Velocity  is  a  vector,  but  speed  is  a  scalar.) 

7.  False.  (The  particle  will  be  at  (2,  -1)  +  2(1,  3)  =  (4,  5).) 

9.  False.  (From  the  parametric  equations,  we  may  read  a  vec- 
tor parallel  to  the  line  to  be  (—2,  4,  0).  This  vector  is  not 
parallel  to  (-2,  4,  7).) 

1 1 .  False.  (The  line  has  symmetric  form 


13.  False.  (The  parametric  equations  describe  a  semicircle 

because  of  the  restriction  on  t.) 

15.  False.  (||*a||  =  |&|||a||.) 

1 7.  False.  (Let  a  =  b  =  i,  and  c  =  j.) 

19.  True 

21 .  True.  (Check  that  each  point  satisfies  the  equation.) 

23.  False.  (The  product  BA  is  not  defined.) 

25.  False.  (det(2A)  =  2"  det  A.) 

27.  False.  (The  surface  with  equation  p  =  4  cos  cp  is  a  sphere.) 

29.  True 


Answers  to  Selected  Exercises  569 


Miscellaneous  Exercises  for  Chapter  1 

3.  x  =  63t  +  1,  y  =  148f,  z  =  847r  -  2 
5.  (a)  x  =  5t  +  2,  y  =  3t  -  2 


(b)  x  =  (b2  -  a2)t  +  (en  +  bi)/2, 
(a2  +  b2)/2 

7.  (a)  2xi  +  4x2  —  2x3  +  X4  —  x5  =  5 
(b)  (bi  -  a\)x\  -\  \-(b„-  a„)x„ 


y  =  (a\  -  bi)t  + 


■a?  + 


9.  (a)  No 

11.  (a)  tt/3 
(b)  x  =  t,  y 


(b)  No 


\-t,z  =  t 


13.  Planes  (a)  and  (d)  are  parallel;  (b)  and  (e)  are  the  same; 
(c)  is  perpendicular  to  (b). 

1 5.  The  dot  product  measures  the  agreement  of  the  answers. 

17.  Hint:  a  x  b  is  normal  to  the  plane  determined  by  a 
and  b. 

19.  (a)  5^21  (b)  10 

21.  (b)  1 

2 5 .  Hint :  Note  that  to  determine  x,  it  suffices  to  determine  ||x|| 
and  the  angle  between  a  and  x.  Consider  the  cases  c  =  0 
and  c  /  0  separately. 

29.  (a)  Hint:  Let  a  and  b  denote  two  adjacent  sides  of  the  par- 
allelogram. 

(b)  Hint:  You  should  simply  give  a  vector  equation. 


31.  (b)A" 


1  n 
0  1 


33.  (a)  H6 


6      7      8      9  10 


J_ 

10 


11 


1  1 

det#2  =  — ,  detff)  =  , 

12  2160 

1  1 

det#4  =  ,  det//5  =  , 

6048000  266716800000 

1 

det  H6  =  

186313420339200000 

(b)  det  #10  «  2.16418  x  10"53 

(c)  Most  likely,  AB  and  BA  will  not  equal 


Sa  +  b 

35.  x  =  (a  +  b)cost  —  bcos  (  — ; — t 


,'a  +  b 

y  =  (a  +  b)  sin  t  —  b  sin  (  — ; —  t 


' a  —  b 

37.  Hypotrochoid:  x  =  (a  —  £>)cos/  +  ccos  |  — : — t 


'  a  —  b 

y  =  (a  —  b)  sin  t  —  c  sin  — ; — t  \  ; 


'a  +  b 

Epitrochoid:    x  =  (a  +  b)  cos  t  —  c  cos  |  — : — t 


,'a  +  b 

y  =  (a  +  b)  sin  t  —  c  sin  |  — : — t 


39.  (a) 


-1.00 


-1.00-0.75-0.50-0.25     0     0.25  0.50  0.75  1.00 


(b) 


570       Answers  to  Selected  Exercises 


Answers  to  Selected  Exercises  571 


(c)  z 


(d)  z 


45.  (a)  {(r,  6,  z)  |  0  <  r  <  3,  0  <  6  <  2n,  0  <  z  <  3} 

if  the  cylinder  is  positioned  so  that  the  center  of  the 
bottom  is  at  the  origin  and  its  axis  is  the  z-axis. 

(b)  Using  the  same  positioning  as  in  part  (a),  {(p,  cp,  6)  \ 
0  <  p  <  3  sec  cp,  0  <  (p  <  jt/4,  0  <  6  <  2n}  U 
{(p,  <p,  6)  |  0  <  p  <  3csc?),  7r/4  <  <p  <  jt/2,  0  < 

e  <  in}. 

Chapter  2 

Section  2.1 

1 .  (a)  Domain  =  R;  range  =  {>>  |  y  >  1} 

(b)  No 

(c)  No 

3.  Domain  =  [(x,  y)  \  y  /  0};  range  =  R 
5.  Domain  =  R3;  range  =  [0,  oo) 
7.  Domain  =  {(x,  y)  \  y  ^  1}; 

range  =  {(x,  y,  z)  |  y  /  0,  y2z  =  (xy  -  y  -  l)2 

+  0+1)2} 
9.  y,  z,  t)  =  xyzt, 

v2(x,  y,  z,  t)  =  x2  -  y2, 

Vi(x,  y,  z,  t)  =  3z  +  t 


11.  (a)  f(x)  =  -2x/W 

(b)  Mx,  y,  z)  =  -2x/Jx2  +  y2  +  z2, 
f2(x,  y,  z)  =  -2y/Jx2  +  y2  +  z2, 
Mx,y,z)  =  -2z/Jx2  +  y2  +  z2 
13.  (a)  fi(\)  =  2xi  -  *3  -  x4, 
/2(x)  =  3x2, 
/3(x)  =  2xi  -X]-x4 
(b)  Range  =  {(yu  y2,  y3)\yi  =  >'3} 


/  — y 

A 

17.  y 


X 


-10      -5        0        5  10 

Z 


x 


Answers  to  Selected  Exercises  573 


574       Answers  to  Selected  Exercises 


45. 


Section  2.2 

1 .  Open 

5.  Neither 

9.  Does  not  exist 

13.  0  15.  0 


3.  Neither 
7.  2 

1 1 .  Does  not  exist 
17.  0  19.  -1 


Answers  to  Selected  Exercises  575 


21.  Limit  does  not  exist. 
23.  Limit  does  not  exist. 
25.  Limit  does  not  exist. 


27.  Limit  is  0. 


11 


9.  fx(x,  y)  =  ey  +  2xy  cos  (x2  +  y), 

fy(x,  y)  =  xey  +  sm(x2  +  y)  +  y  cos(x2  +  y) 
Fz  =  V(y  +  z), 


F- 


:  -(x  +  z)/{y  +  zf, 
(y-*)/(y  +  z)2 


13.  Fx  =  x/^Jx2  +  y2  +  z2, 
Fy  =  y/Jx2  +  y2  +  z2, 
Fz  =  z/^Jx2  +  y2  +  z2 

15.  F 


1  -  2x2  -  3xy  +  y2  -  3xz  +  z2 


Fy 


17.  Fx 

Fy 


(l+x2+y2  +  z2)V2 

1  +  x2  -  3xy  -  2y2  -  3yz  +  z2 
{\  +  x2  +  y2  +  z2?'2 

1  +  x2  +  y2  -  3xz  -  3yz  -  2z2 
(1  +  x2  +  y2  +  z2)5'2 

x4  -  2xyz  +  3x2z2  +  3x2 


(x2  +  z2+  l)2 


19. 
21. 
23. 


F- 


x2  +  z2  +  r 

x2y  —  2xiz  —  yz2  +  y 
(x2  +  z2  +  l)2 


-i  +  (27T  +  l)j 

+  j-2k 


2k 


25.  i 

29. 

Limit  does  not  exist. 

27.  [-4  -3/e 

-3/e] 

31. 

2 

29. 

0  -2 

0 

33. 

Limit  does  not  exist. 

_  1/V5  0 

-2/V5 

35. 

0 

_ 

37. 

Limit  does  not  exist. 

"  3  -7 

1 

0  " 

39. 

Continuous 

31. 

5  0 

2 

-8 

41. 

Continuous 

_  0      1  - 

17 

3  _ 

43. 

Not  continuous  at  (0,  0) 

"  -2      0  " 

45. 

Continuous 

33. 

1  -1 

47. 

Hint:  Write  /(x)  in  terms  of  the  components  of  x. 

0  2 

Section  2.3 

1 .  fx(x,  y)  =  y2  +  2xy,  fy(x,  y)  =  2xy  +  x2 
3.  fx(x,  y)  =  y  cosxy  —  y  sinxy, 

fy(x,  y)  =  x  cosxy  —  x  sinjcy 
5.  fx(x,y)  =  4xy2/(x2  +  y2)2, 

fy(x,  y)  =  -4x2y/(x2  +  y2)2 
7.  fx(x,  y)  =  —  3x2y  sinx3y, 

fy(x,  y)  =  —x3  sinx3y 


37.  (a)  The  function  has  continuous  partial  derivatives. 

(b)  z  =  3x  +  Sy  +  3 
39.  z  =  e(x  +  y) 

41 .  x5  =  -4x\  +  6x2  -  4x3  -  6x4  +  28 
43.  (a)  h(0.l,  -0.1)=  1 

(b)  /(0.1,-0.1)=  1 
45.  (a)  fe(1.01,  1.95,  2.2)  =  21.76 

(b)  /(1.01,  1.95,  2.2)  =  21.6657 
47.  (a)  fx(x,y)  =  (3x4  +  Sx2y2  +  y4)/(x2  +  y2)2, 

f  (x,  y)  =  -{x4  +  4x2y  +  2x2y2  +  y4)/(x2  +  y2)2 


(b)  fx(0,  0)  =  3,  /,(0,  0) 


■1 


576       Answers  to  Selected  Exercises 


49.  (a)  (8  +  A4l)x  -  8y  =  V2(n  -  4) 
(b) 


0  0.5 

51.  (a)  x  +  y  =  ln2-  1 
(b) 


53.  (a)  z=Ux-  15 
(b) 


2.50 


0.50 


(c)  Partial  derivatives  are  polynomials — hence  contin- 
uous— so  /  is  differentiable  at  (2,  1). 


(c)  Partial  derivatives  are  rational  functions  defined  on 
R2 — hence,  continuous — so  /  is  differentiable  at 
(0,  0). 

57.  (a)  z  =  -j^[(9-j3n  -  96V2>  -  (16V2tt  +  72)y  + 
(4V2  -  3V3)tt2  +  (1672  +  9)tt] 

(b) 


(c)  Partial  derivatives  are  products  of  polynomials  and 
sine  and  cosine  functions  and,  hence,  are  continuous. 
Thus,  /  is  differentiable  at  (tt/3,  Jr/4). 
59.  Df(x)  =  A.  Note  that  f'(x)  =  a  in  the  one-variable  case. 

Section  2.4 

1-  D(f  +  g) 

=  [y  —  sinx  +  y  cos(xy)       x  +  3y2  +  x  cos(xy)] 
3.  D(f+g) 

siny  +  3x2  cos x  —  x3  sinx      xcosy  1 

—6x  +  yz  ez+xz      yez  +  xy 

5.  D(/g)=[3x2  +  y2  2xy], 

D(f/g)=[y2-y4/x2       2xy  +  4y3/x] 
7.  D(fg) 

=  [l2x3y  +  3x2y5  -  12xy3  -  2y7 

3x4  +  5x3y4  -  18x2y2  -  14xy6], 
D{f/g) 

~y(2y6  -  6x3  -  3x2y4)  3x3  +  6xy2  +  5x2y4  -  6y6 
x2(x2  -  2y2)2  x(2y2  -  x2)2  , 

9.  /„  =  6xy\  f„  =  42x3y5  +  6x, 
fxy      /,,     2\vy<  ■  6  v  7 
11.  fx. 
fx. 

13.  /, 


-ye~x  +  2yx-ie>'x  +  ylx~"eyix ,  fyy 
fyx  =  e~ 

2[sin2  2x  —  cos  2x(sin2  x  +  2ey)] 


(sin2  x  +  2ey)3 


fyy  = 

15.  /„  = 
siny 


2ey(2ey  -  silt  x) 
(sin2x  +  2ey)3 


fxy  —  fyx 


4ey  sin  2x 
(sin2x  +2ey)3 


-y  sinx,    fyy  =  x  cosy,    fxy  =  fyx  =  cosx  + 


Answers  to  Selected  Exercises 


577 


17.  /„  =  2e\  fyy  =  x2e>,  fzz  =  Ae2\  fxy  =  fyx  =  2xe> ', 

/,,     ./:-.     0,  j\:     ./.,  0 
1 9.  fxx  =  2yz,  fyy  =  2xz,  fa  =  2xy,  fxy  =  fyx  =  2xz  + 

2yz  +  z2,    fxz  =  fzx  =  2xy  +  y2  +  2yz,    fyz  =  fzy  = 

x2  +  2xy  +  2xz 

21 .  fxx  =  b2ebx  cosz  +  a2eax  siny,  fyy  =  —eax  sin  y,  fzz  = 

-ebx  cos  z,       fxy  =  fyx  =  aeax  cos  y,       fxz  =  fzx  = 

-bebxsmz,fyz  =  fzy  =  0 
23.  d"f/dx"  =  3"yelx,  d"f/dy"  =  0,  forn  >  2 
25.  d"f/dx"  =  {-\)n-\n  -  \y./x",d"f/dy"  =  (-l)"-\n  - 

l)!/y",  d"f/dz"  =  (-1)"(«  -  l)l/zn.  All  mixed  partials 

are  zero. 

27.  (a)  px  and  py  have  degree  16;  pxx,  pyy,  and  pxy  all  have 
degree  15. 

(b)  px  and  py  have  degree  3;  pxx  has  degree  2;  pyy  has 
undefined  degree;  pxy  has  degree  2. 

(c)  The  degree  of  dkp/dxil  ■  ■  ■  dxik  is  d  —  k,  where  d 
is  the  highest  degree  of  a  term  of  p  of  the  form 
xf'x*  •  •  ■  xft"  such  that,  for  j  =  1, 2, . . . ,  n,  dj  is  at 
least  the  number  of  times  Xj  occurs  in  the  partial 
derivative.  If  p  has  no  such  term,  then  the  degree  of 
dkp/dxil  ■  ■  ■  dxit  is  undefined. 

29.  (a) 


T  0 


(b) 


Z  0 


33.  (a) 


Section  2.5 

1.  df/dt  =  (18f2  +  14f)sin2?  -  6 cos 2t  sin2  2t  +  (12r3  + 
14r2)cos2r +  72r  +  84 

2.  df/ds  =  (3s2  +  t2+  2st)  cos(0s  +  t)(s2  + 12)), 
df/dt  =  (s2  +  3t2  +  2st)cos((s  +  t)(s2  +  t2)) 

3.  (a)  dP/dt  =  (36  -  277r)/V2  atm/min 
(c)  Approximately  (.36  +  (9  -  .27)tt)/V2 

«  19.6477  atm 
5.  —6  in3/min.  Decreasing. 
7.  0.766  units/month 
9.  0.2244  cm/min 
13.  Hint:  dw/du  =  2u(dw/dx)  —  2u(dw /By)  from  the  chain 
rule. 

15.  Hint:  dw/dx  =  (dw/du)({y^  -  x2y)/(x2  +  y2)2)  from 
the  chain  rule. 

17.  Hint:  dw/dx  =  —\/x2(dw/du  +  dw/dv)  from  the  chain 
rule. 

19.  D(fog)=      2e2j_14/  _14^-i4f 


578       Answers  to  Selected  Exercises 


21.  D(fog)=[(l+s  +  t)es-'  (\-s-t)es->] 

3s2  -t2  -  1st 

6s5P-s-2r2      3s6t2  -25-1/-3 


23.  D(fog) 


25.  D(fog): 
27.  D(fog) 


e2,{2  sinf  +  cosf) 
e'(sin2  t  +  2  sin  t  cos  f) 
3(sin2rcosr  +  e3') 


29.  (a) 
(b) 
31.  (a) 


t  +  u 

3s2f3  -  tu2es'"2 


7  10 
31  44 


S  +  M 


3*3r2 


1  13 

2  31 


dx2 


drL  r 


sin  6  3      2sin#  cosf?  3 


3r 


3r36» 


2  sinfl  cos  9  3  sin2  6»  32 
+  =  —  + 


iil) 


r2  d92' 


By2 


sin2  6 


dr2 


2      cos26  3      2sin#cos#  32 
+  —  H  


3r 


3r36> 


2  sin  6  cos  0  3      cos2#  32 
+ 


33.  (a) 


dr 


r2  36 

3      cos  <p  3 

sin  <p  1  

dp        p  dtp 


r2  d62 


35.  dy /dx  =  (2xy1  —  y  cos xy)/(x  cosjty  —  lx2y6  +  ey) 

37.  dz/dx  =  3x2z3/(yz2  sinz  +  siny  —  x3z2),  dz/dy  = 
(z2  cos  z  +  z  cos  y)/(y z2  sin  z  +  sin  y  —  jc3z2) 

'  dw\  ( dw\  ( dw^ 


39.  (a) 


_  i  (dw\     -  7 

dx  J  y  z       '  \  dy  Jx.z       '  V  3z 

.  „  /  dw\    / dw 


1  -  20y 


(b) 


dw 

~dx 


(dw 
~dx 


'  dw\      /  dy\       ( dw 
dy  Jx.z  \  dx  J       \  dz 


3z\ 


41. 


3  s 
dz 


(l)(l)  +  (7)(0)  +  (-10)(2*) 
=  xw  —  2z; 


xw  —  2z  + 


2   3  3 

x  y  —  x 
xw  —  3y2z 


Section  2.6 

1.  (a)  The  directional  derivative  of  /  at  (x,  y,  z)  in  the 
negative  z-direction. 
(b)  V/(*,y,z)-(-k)  =  -3//3z 
3.  8V5/5 
5.  (2e- 9)75/5 
7.  -4V3e-14 

9.  (a)  fx(0,  0)  =  /,(0,  0)  =  0 

(b)  -Dv/(0,  0)  =  v\w\  for  all  unit  vectors  v  =  v  i  +  w  j. 
(c) 


11.  (a)  (— 2i  +  j)/V5  direction 
(b)  ±(i  +  2j)/V5  direction 

13. 


1  5.  Travel  along  y3  =  27 x  (toward  the  origin) 

17.  y-z  =  \  19.  x-2y-2z  =  2 

21 .  The  tangent  plane  has  equation  x  —  nz  =  2. 

23.  (|, -£,_§)  „d  (-f.s,  9) 

27.  (a)  3jc+12v  +  2z+10  =  0 

(b)  There  is  no  tangent  plane  at  (0,  0,  0). 
29.  Tangent  line  is  5x  —  3^/4  y  =  —  1. 
31 .  Parametric  equations:  x  =  5t  +  5,  y  =  4t  —  4.  Cartesian 

equation:  Ax  —  5y  =  40. 
33.  Parametric  equations:  x  =  14f  +  2,  y  =  t  —  1.  Cartesian 

equation:  x  —  14y  =  16. 


Answers  to  Selected  Exercises  579 


35.  x  =  -2t  -  1,  y  =  t  -  1,  z  =  t  -  1 

37.  Xs  —  X\  =  71 

39.  xi  +  x%  H  1-  xB_i  —  x„  =  V« 

41 .  (a)  Near  all  points  of  S  such  that  x  /  0.  At  such  points 
f(x,  y)  =  ln(l  —  sinxy  —  x3y)/x. 
(b)  All  points  (0,  y,  z)  (i.e.,  the  yz-plane). 
(c) 


43.  (b)  Yes,  y  =  x2?\ 


y 


1  - 

0.5- 

-0.5 

0 

0.5 

X 


(c)  The  implicit  function  theorem  suggests  that  we  need 
not  be  able  to  solve  for  y  in  terms  of  x .  Note  that  we 
can  in  this  case,  but  that  the  function  fails  to  be  of 
class  C1  at  x  =  0. 

45.  (a)  G(-l,  1,  1)=  F(-2,  1)  =  0 

(b)  Hint:  Use  the  chain  rule  to  find  Gz(-l,  1,1)  = 
30/0. 

47.  (a)  A(l,  0,-1,  1,2)  =  -120/0;  apply  the  general  im- 
plicit function  theorem. 

(b)  ^(1,  0)  =  -I  j^(l,  0)  =  -|,  ^(1,  0)  =  -1 
oxi  oxi  axi 

49.  (a)  Anywhere  except  where  <p  =  0  or  it . 

(b)  We  can  solve  anywhere  except  along  the  z-axis,  which 
is  where  6  can  have  any  value. 


Section  2.7 

1.  (1.302942538,-0.902880451) 
3   (a)  Estimate  intersection  points  near  (1,  1/2)  and 
(-1/2,  -3/4). 


y 


1.5 

- 

1.0 

0.5 

-1.J 

-1.0 

-0.5 

-0.5 

0.5 

i 

-1.0 

-1,5 

(b)  (1.103931711,0.441965716)  and  (-0.518214315, 
-0.657923613) 
5  (a)  x0  =  (-1,  1)  leads  to  (-1.26491 11,  1.54919334). 

(b)  x0  =  (l,-l)  leads  to  (1.26491106,-1.5491933), 
while  x0  =  (-l,-l)  leads  to  (-1.2649111, 
-1.5491933). 

(c)  It  appears  that  an  initial  vector  leads  to  an  intersection 
point  that  lies  in  the  same  quadrant. 

9.  (-0.9070154,  -0.9070154) 


True/False  Exercises  for  Chapter  2 

1 .  False 

3.  False.  (The  range  also  requires  v  /  0.) 

5.  True 

7.  False.  (The  graph  of  x2  +  y2  +  z2  =  0  is  a  single  point.) 

9.  False 

1 1 .  False,  (lim^^-^o.o)  f(x,  y)  =  0  /  2.) 

13.  False 

1 5.  False.  (V/(x,  y,  z)  =  (0,  cos  y,  0).) 

17.  True 

1 9.  False.  (The  partial  derivatives  must  be  continuous.) 

21.  False.  (./,,  .-  /,,.) 

23.  True.  (Write  the  chain  rule  for  this  situation.) 

25.  False.  (The  correct  equation  is  x  +  y  +  2z  =  2.) 

27.  True  29.  False 


580       Answers  to  Selected  Exercises 


Miscellaneous  Exercises  for  Chapter  2 

1  .(a)  /i(x)  =  -x2,  fi(s)  =  x\-  *3,  /3(x)  =  x2 
(b)  Domain  is  R3 ;  range  consists  of  all  vectors  in  R3  of  the 
form  a  i  +  b  j  —  a  k. 
3.  (a)  Domain:    {(x,  y)  \x  >  0,  y  >  0}  U  {(x,  y)  |  x  <  0, 
y  <0} 

Range:  [0,  oo) 


(b)  Domain  is  closed. 


5. 


f(x,y)  = 

Graph 

Level  curves 

1 

x2  +  y2  +  1 

D 

d 

sin-y/x2  +  y2 

B 

e 

(3y2  -  2x2)e-x2-2y2 

A 

b 

y3  —  3x2y 

E 

c 

x2y2e-S-2f 

F 

a 

-v2-v2 

ye  x  } 

C 

f 

7.  0 

11.  (a)  15°F 
(b)  5°F 

13.  1.3 

15.  (a)  9.57°F;  16.87°F;  8.39mph 

(b)  Effect  of  windspeed  is  greater  in  the  Siple  formula 
than  in  the  table. 

(c)  If  t  <  91.4°F  or  s  >  4  mph,  the  Siple  formula  gives 
windchill  values  that  are  higher  than  the  air  tempera- 
ture, which  is  unrealistic. 


120 


(b) 


w  (io, o 


20 

,  ,  yf^r     ,  ,  i 

-40  -20 

^20 

40 

-40 

-60 

(c) 


Wi 


Answers  to  Selected  Exerc' 


Wn 


1 9.  Hint:  Compare  the  normal  vector  to  the  tangent  plane  with 

the  vector  O  P . 
21.  (a)  3*  -  4y  -  5z  =  0 
(b)  ax  +  by  —  cz  =  0 
23.  (a)  2x  -  8y  -  z  =  5 

(b)  x=l,y  =  t-l,z  =  5-8t 
25.  12,  201.4  units/month 
27.  (a)  dw/dp  =  2p,  dw/dip  =  dw/98  =  0 
29.  (a)  dz/dr  =  e'(cos  6(dz/dx)  +  sin 0(dz/dy)), 
dz/96  =  er(-  sme(dz/dx)  +  cos6(dz/dy)); 
dz/dx  =  e-r(cos6(dz/dr)  -  sm0(dz/d0)), 
dz/dy  =  e-r{sm6(dz/dr)  +  cos8(dz/d6)) 
(b)  Hint:  d2z/dx2  =  e-2r(cos2  6(d2z/dr2) 

+  (sin20  -  cos2  6)(dz/dr) 
+  2cos6»sin6»(9z/96») 
-2cos6»sin6»(92z/9r90) 
+  sin26»(92z/902)), 

-2'(sin26l(92z/9r2) 


d2z/dy2 


+  (cos26»  -  sin26»)(9z/9r) 
-2sin6»cos6»(9z/96») 
+  2sin6»cos6»(92z/9r96l) 
+  cos2  6(d2z/d62)) 

31.  u"  [u{""-l)  +  uu"  Inu  +  u""(\nu)2] 

35.  (a)  z  =  f{x,  y),   where   fix,  y)  =  (x3  -  3xy2)/(x2  + 
y2)  ifix,  y)  /  (0,  0);  fix,  y)  =  0  if  (*,  y)  =  (0,  0) 

(b)  Yes 

(c)  df/dx  =  (x4  +  6x2y2  -  3y4)/(x2  +  y2)2, 
df/dy  =  -8x3y/(x2  +  y2)2  (if  (x,  y)  /  (0,  0)); 

MO,  0)  =  i,  fyio,  o)  =  o 

(d)  £>u/(0,  0)  =  (3//9r)Uo  =  cos  39 

(e)  /y(0, 0)  =  0     from     (c),     but     along     y  =  x, 
fyix,  x)  =  —2,  which  does  not  have  a  limit  of  0. 


37.  Homogeneous  of  degree  3 
39.  Homogeneous  of  degree  3 
41 .  Homogeneous  of  degree  0 

Chapter  3 
Section  3.1 


7.  \(t)  =  3i  +  2j,  speed  =  Vl3,  a(f)  =  0 


582       Answers  to  Selected  Exercises 


9.  v(f)  =  (sin?  +  t  cost,  cost  —  t  sin?,  2t), 
speed  =  V5f2  +  1, 

a(f )  =  (2  cos t  —  t  sin t ,  —2  sin t  —  t  cost,  2) 
11.  (a) 


15.  I(f)  =  fi  +  (3/  +  l)j 

17.  1(f)  =  {At  -  4,  \2t  -  16,  80r  -  128) 

19.  (a)  y 


(b)  1(f)  =  (t,  10?  -  15) 


(c)  y  =x3  -2x+\ 

(d)  y  —  5  =  10(x  —  2).  (Let  x  =  t  to  verify  agreement.) 
21.  117.1875  ft 

23.  25.09°  or  64.91° 
25.  No 

27.  Write  out  both  sides  in  terms  of  component  functions. 
29.  Hint:  Differentiate  ||x(r)||2. 


33.  (a)x(l)  =  x(-l)  =  (l,0) 
(b)  jt/2 


Answers  to  Selected  Exercises  583 


Section  3.2 


0.5    1    1.5    2    2.5  3 


(b)  The  path  consists  of  three  C 1  pieces:  one  for —2  <  t  < 
0,  another  for  0  <  t  <  1,  and  another  for  1  <  t  <  2. 

(c)  4V2 

1 5.  Hint:  Begin  with  x  =  r  cos  6,  y  =  r  sin£>. 

(-5  sin  3f  i  +  2j  +  5  cos  3f  k), 


17.  T: 

N  : 

B 

K  - 
19.  T: 

N  : 

B 

K  - 


29 
cos3ri 


sin3f  k, 


1 


(-2sin3f  i-  5j  +  2cos3fk), 


29'  1 

2 
3 


'  24 


(1,  ±Vf+T, -±VT=f), 

2V2(o,  iyr^r,  ivm"), 


■^(l,-Vt  +  T,VT=t), 

V2/(67T^72),  r  =  1/  (sVT^T2  ) 


21.  (b)  K 
23.  (a) 


sin  x 


ab 


25.  (a) 


(a2  sin  t  +  b2  cos2  tfl1 


25 
20 
15 
10 
5 
0 


584       Answers  to  Selected  Exercises 


27.  atang  =  4f/V4f2  +  1,  anorm  =  2/V4f2  +  1 

29.  atang  =  V5e',  flnorm  =  2a/5V 

31 .  atang  =  4f/V2  +  4f2,  anorm  =  2/Vl  +2f2 

33.  (b)  0^  =  4f/V4f2  +  10,  anorm  =  27T0/V4f2  +  10 

35.  Hint:  Write  v  and  a  in  terms  of  T  and  N  and  use  the 

Frenet-Serret  formulas. 
37.  Hint:  Use  formula  (17)  and  Exercise  35. 
39.  Hint:  Recall  that  ||x  —  xo||2  =  (x  —  xo)  •  (x  —  xo). 
41 .  Hint:  Calculate  T  and  show  that  (x(r )  -  (1 ,  0,  0))  •  T  =  0. 

Section  3.3 

1.  y 


9. 


/  /  /  /  -  - 

/  /  /  y  '  - 

///'■■- 

V  V  \  \ 

////,. 

III!.. 
Ill  

v  \  V  \ 
»    >    \  V 

\      \      \      V      x  - 

\    \    \    V    x  « 

i    i    i   i  i 
'    '    /  / 
'   '   /  1 

\  \  N   V    -  - 

'  '  /  / 

\\  N      -  - 

'  s  /  / 

- 

/  /  /  / 

//// 

///  / 

/  1 

\  \ 

\  W\ 

// /  f 

/  I 

»  \ 

\  \  w 

/  /  /  ' 

/  i 

\  \ 

/  s  /  / 

/  / 

S   \  X  X 

>k  V  v  s 

N  V  \  \ 

\  \ 

/  / 

✓  y  s  s 
/  /  /  S 

\\  \  \ 

\  V 

/  / 

/  /  /  / 

\\  \  \ 

\  \ 

;  / 

/  /  // 

\\\  \ 

\  \ 

/  / 

/  /// 

11. 


13. 


.    .  -        X  "V  \ 

\  \  \  \ 
\  v  \  \ 
y  \  \  t 

1  t  t 
t  t  f 
t  f  / 

„_-.---••  s 

V     \     t  t 

f  f  / 
t  I  1 
1  f  t 

v    i    t  » 

k  1  i  y 

k  1   '   '  ' 

-    t  J 

/  /  / 

4    t  / 

///'''' 
///>'''• 
/   /    /    t    f    '  ' 
/    i    t    *    ♦    '  ' 

/  M  1  '  n 

J  (  M  V  *  > 

x  x  ^  ^ 

1  1  \  \  \  v  \ 

Answers  to  Selected  Exercises  585 


15. 


Section  3.4 


■  •  i 

4  4  *■  » 


f  t  '  ' 
t  1  ' 


~  4  * 
~  4  4 
-  <  < 


'  '  t 


17.  F(x(f))  =  (cosf,  -sinf,0)  =  x'(0 


19.  F(x(f))  =  (cosf,  -sin?,2e2')  =  x'(0 


21. 

23. 

25. 
31. 


1 


2-? 

(a)  F  =  V/,  where  f(x,  y,  z)  =  3x  —  2y  +  z. 

(b)  Equipotential  surfaces  are  parallel  planes  with  equa- 
tion 3x  —  2y  +  z  =  c  (i.e.,  planes  with  normal 
(3,  -2,  1)). 

Hint:  Use  the  chain  rule  and  the  facts  that  V/  =  F  and  x 
is  a  flow  line  of  F. 

Hint:  Differentiate  the  defining  differential  equation  for 
the  flow  with  respect  to  x  =  (*i,  X2,  ■  ■  ■ ,  x„). 


1 
3 
5 
7 
9 
11 
13 


2yz)k 


div   F  <  0 


div   F  <  0  on 


2x  +  2y 
3 

2x\  +  4x2  +  6x3  +  •  ■  ■  +  2nxn 
2xz  i  —  2yz  j  —  ey  k 
0 

(x2  -  xyexyz)i  +  (y2  -  2xy)\  +  (yzexyz 

(a)  div  F  >  0  on  all  of  R2 

(b)  divF  <  0  on  all  of  R2 

(c)  div   F  >  0   on   {(x,  y)\x  >  0}, 

{(*,  y)  \x  <  0} 

(d)  div  F  >  0   on   {(x,  y)  |  v  <  0}, 

{(*,  y)  I  y  >  0} 

21 .  Write  out  in  terms  of  the  component  functions  of  F  and  G. 
23.  Write  out  in  terms  of  the  component  functions  of  F. 
25.  Write  out  in  terms  of  the  component  functions  of  F  and  G. 
27.  First  use  the  chain  rule  to  replace  occurrences  of  the 

Cartesian  differential  operators  in  V  by  combinations 

of  spherical  differential  operators.  Then  compute  V/. 

Finally,  replace  i,  j,  and  k  by  appropriate  combinations 

of  ep,  e^,  and  eg.  (See  also  §1.7.) 
29.  Write  out  the  components  of  /V/. 
33.  (1/V3.4/V3.-1/V3) 


True/False  Exercises  for  Chapter  3 

1 .  True 

3.  True 

5.  False.  (There  should  be  a  negative  sign  in  the  second  term 
on  the  right.) 

7.  True 

9.  False 

1 1 .  True 

13.  True 

1 5.  False.  (It's  a  scalar  field.) 

17.  True 

1 9.  False.  (It's  a  meaningless  expression.) 

21 .  True.  (Check  that  F(x(0)  =  x'(O-) 

23.  False.  (V  x  F  /  0.) 

25.  False.  (Consider  F  =  y  i  +  x  j.) 

27.  True 

29.  False.  (V  •  (V  x  F)  /  0.) 


Miscellaneous  Exercises  for  Chapter  3 


1.  (a)  D       (b)  F       (c)  A 
(d)  B       (e)  C       (f)  E 

3.  Hint:  Differentiate  (ds/dt)2  = 


x'(0 


586       Answers  to  Selected  Exercises 


w  =  0 


9.  (a)  Hint:  Calculate  x(0)  and  x(l). 
(b)  Hint:  Show  that 


11  ■  (a)  mwj^  ~  2x2  +  -*3)2  +  (vi  -  2y2  +  y3f 


(b) 


X] 


2x2  +  x3)2  +  (yi  -2y2  +  y3)2 


2(l+u.)^ 

13.  (a)  Hint:  Find  where  x'(f)  =  (0,  0). 

(b)  Hint:  The  tangent  line  at  x(to)  is  given  by  \(s)  : 
x(?o)  +  Jx'(?o).  Find  the  y-intercept  of  this  line. 
15.  Hint:  Begin  with  x(6>)  =  (/(0)cos6>,  f(6)sm0). 
17.  (a)  y(f)  =  (a(cos  f  +  t  sinr),  a(sinf  —  t  cos  f)) 

(b)  y 

4 


19.  (a)  Show  that  ||x(r)  -  y(r)||  = 

(b)  The  vector  difference  x(f)  —  y(t)  is  a  tangent  vector 
(to  x)  s(t)  units  long. 
23.  e(f)  =  (a(?  +  sin;),  a(cosf  —  1)),  which  is  a  another  type 

of  cycloid. 

25.  Hint:  Use  the  Frenet-Serret  formulas. 
27.  (a)  Show  that  (x'(s))2  +  (y'(s))2  =  1  by  means  of  the 
fundamental  theorem  of  calculus. 


(b)  K  =  \g'{s)\ 

(c)  Set  g'(s)  =  k(s).  Find  g  by  antidifferentiation  and  set 
x(s)  =  So  cos g(t)dt,y(s)  =  So  sin g(t)dt. 

(d)  x(s)  =  f*  cos(t2/2)dt,  y(s)  =  f°sm(t2/2)dt 
(e) 


29.  (a)  Calculate  B  and  remember  that  it  is  a  unit  vector, 
(b)  B  is  constant,  so  B'  =  0.  Next  use  the  Frenet-Serret 
formula. 

a2  +  h2/4jt2 

31.  (a)   (b)  8.2771  ft 

a 

35.  N'  =  -kT  +  tB.  Argue  that  N'  /  0. 

37.  (a)  242tx        (b)  F  =  V2(y  i  -  2x  j  +  y  k) 

41 .  Hint:  Use  Proposition  1.4. 

43.  No 


Chapter  4 


D3  + 


9)3/3888 


Section  4.1 

p4(x)  =l+2x  +  2x2  +  4x^/3  +  2x4/3 
p4(x)  =  1  -  2(x  -  1)  +  3(x  -  l)2  -  4(x 
5(*  -  l)4 

P3(x)  =  3  +  (x  -  9)/6  -  (x  -  9)2/216  +  (x 
Ps{x)  =\-{x-  nil)2 12  +  (x  -  7r/2)4/24 
Pi{x,  y)=\-2{x-  l)/9  +  2(y  +  l)/9, 

Pz(x,  y)=\-2(x-  l)/9  +  2(y  +  l)/9  +  (x  -  l)2/27 
-  8(x  -  l)(y  +  l)/27  +  (y  +  l)2/27 
Pi(x,  y)  =  —  1  —  2x, 

p2(x,  y)  =  -l-  2x  -  2x2  +  9(y  -  tt)2/2 


1. 
3. 

5. 
7. 
9. 


11 


13.  pi{x,  y,  z)  ■ 

3y2  + 2xz 
15.  p\(x,  y,  z)  - 


1  +  x  +  Sy  +  4z 


p2(x,  y,z)  =  0 


p2(x,  y,  z)  =  xy 


17. 


VI 

4 
4 


v2 

4 

4  -I 


Answers  to  Selected  Exercises  587 


19. 


6  2  0 
2  0-2 
0   -2  12 


21.  P2(*,y)  =  l  +  [  0  0  ] 
23.  p2(x,y,z)  =  2+[  0   5  1] 


2  0 
0  -2 


z-2 


"  0 

3 

0 

X 

jc   y   z  —  2  ] 

3 

8 

2 

y 

0 

2 

0 

_  z-2  _ 

25.  (a)  Df(0)=[  1   2   •••   n  ], 
"  1    2     3  ••• 


ff/(0) 


2  4  6 

3  6  9 


2n 

3/7 


«   2h    3«    •  •  •  «2 

(b)  Pi(*i,  .  .  .  ,  Xn)  =  1  +  Jti  +  2x2  H  1"  «*n 

p2(*l,  .  .  .  ,  A„)  =  1  +  £"=1  «i  +  \  E"  ,=i  ?7*<*/ 

27.  (a)  2  -  z  +  3xy  +  x2y  -  xz2  +  2y3 

(b)  -4  -  4(x  -  1)  +  ll(y  +  1)  -  z  +  |[4(jc  -  l)2  + 

16(*  -  l)(y  +  1)  -  12(y  +  l)2  -  2z2]  + 

1[18(jc  -  l)3  +  24(x  -  l)2(y  +  1)  -  6(x  -  l)z2  + 

12(y  +  l)3] 

29.  2x  dx  +  6y  dy  —  6z2dz 

31.  ex  cos  j      +  (e-v  sin z  —  eA'  sin y)dy  +  ey  cos z dz 
33.  (a)  388.08       (b)  0.24625       (c)  1.1 
35.  The  ( 1 ,  l)-entry  (upper  left) 
37.  2.4  cm 
39.  0.0068  m 

41.  (a)  P2(x,y)=\-\[x2+{y-«1)2} 
(b)  Accurate  to  at  least  0.0360 

43.  (a)  p2(a.,y)=|_y_2x(y-f) 
(b)  Accurate  to  at  least  0.03 1 1 

Section  4.2 

1.  (a)  (2,3) 

(b)  Af  =  -h2  -  k2 

(c)  There  is  a  local  maximum  at  (2,  3). 


3.  Local  maximum  at  (|,  |) 

5.  Local  minimum  at  (y,  5);  saddle  point  at  (|,  l) 

7.  Minimum  at  (4,  |) 

9.  Saddle  point  at  (0,  0);  minimum  at  (0,  2) 

1 1 .  Saddle  point  at  (0,  ^);  local  minimum  at  (0,  —  ) 

1 3.  Local  maximum  at  (— j,  |) 

15.  Saddle  point  at  (0,  6,  -3) 

17.  Local  minimum  at  (0,  0,  0) 

19.  Saddle  point  at  (-1,  \,  \) 

21.  (a)  (0,  -2)  and  (0,3) 

(b)  Local  maximum  at  (0,  —2); 

local  minimum  at  (0,  3) 

23.  (a)  Minimum  if  a,  b  >  0;  maximum  if  a,  b  <  0;  saddle 
point  otherwise 

(b)  Minimum  if  a,  b,  c  >  0;  maximum  if  a,  b,  c  <  0; 
saddle  point  otherwise 

(c)  Minimum  if  a\, . . . ,  an  >  0;  maximum  if  a\, . . ., 
an  <  0;  saddle  point  otherwise 

25.  Saddle  points  at  (0,0),  ^±-^,0^;  local  maxima  at 

(tt  (~75'  7s) 

27.  Saddle  points  at  (0,  0,  0,  0),  (-V2,  2V2,  1 ,  -V2), 
(V2,  2V2,  -1,  — n/2)  {-42,  -2V2,  -1,  4%, 
(a/2,  -2V2,  1,a/2) 

99    ^36   _48  _m 
V 13  '      13'  13,1 

31 .  1 100  units  of  model  X  and  700  units  of  model  Y 
33.  Maximum  of  8  at  (0,  0,  2); 

minimum  of  —  ^  at  (—  y ,  3,  |) 
35.  Maximum  of  1  at  (7r/2,  0),  (jt/2,  2tt),  (3tt/2,  n); 

minimum  of  —  1  at  (3jt/2,  0),  (3n/2,  2n),  (jt/2,  tt) 
37.  Maximum  of  11  at  (2,  0);  minimum  of  |  at  (i,  1 ) 

39.  Maximum  of  e6  at  (0,  1,  —2);  minimum  of  eat  all  (x,  y,  z) 

such  that  x2  +  y2  -  2y  +  z2  +  4z  =  0 
41.  (b)  Maximum 
43.  (b)  Neither 
45.  (b)  Maximum 

47.  (a)  Local  maximum  at  (0,  0,  0) 

(b)  /(0,  0,  0)  =  e2  is  a  global  maximum. 

49.  Global  minimum  of  2  +  ln2  at  (2,\).  No  global 
maximum. 

51.  (a)  {(1,  y)\y  eR}U{(*,2)|*  eR) 

(b)  Maxima  of  3  along  critical  points  in  (a). 

53.  (b)  Critical  points  are  (2,  1)  and  (0,  —1). 


588       Answers  to  Selected  Exercises 


(c) 


z 


Section  4.3 

1 .  (a)  Minimize   f(x,  y)  =  x2  +  y2  +  (2x  -  3y  -  A)2  to 

find  that  the  closest  point  is  (=,—  =  ,  —  =). 

(b)  Minimize  f(x,  y,  z)  =  x2  +  y2  +  z2  subject  to  2x  — 
3y-z=4. 

3.  (V2,  72)  and  (-^2,  -72) 

5.  (1,  |,  2),  (3,  0,0),  (0,2,0)  and  (0,0,  6) 

'  •  V 1 1 '  n  '  11/ 

11-  (±7C7r.^)"d(±^.-A.-^) 

13.  (a)  (±yf,>),(0,±^) 

(b)  (=t-y/|>      give  maxima;  ^0,  ±^j)  give  minima. 

17.  (±1,0,0),(0,±1,0),(0,0,±1),  (|,-|,  j), 
(_2  _2  _n 

V     3'      3'  3/' 

( 1  /H  —1  _i  /TT \  /   i  /TT  _3  3  /TT\ 

^8V  2  '      8'      S\  2  )'\     Sy  2  '      8'  8V  2  J 

19    / J_  _L   krvg   i-V2\  _j_   1+yg  1+VI\ 

'7-  v/2'      2     '      2  VS'      V2'      2     '      2  J 

21 .  The  numbers  are  6,  6,  6. 

23.  Maximum  value:  6.  Minimum  value:  0. 

25.  Height  should  be  equal  to  diameter. 

27.  Locate  at  either  (-2,  2,  1)  or  (-2,  -2,  1). 

29.  Largest  sphere  has  equation  x2  +  y2  +  z2  =  2. 

31.  (|,2,  f) 

33.  Highest  point  is  (— 1,  —  1,  2);  lowest  point  is  [j,  i,  |). 


35.  (a)  (76,  76)  and  (-76,  -76) 

39.  Nearest  points  are  (75,  —  75)  and  (— 75,  75); 

farthest  points  are  (5,  5)  and  (—5,  —5). 
41.  (a)  Critical  point  at  (1,0) 

(b)  There  is  a  minimum  of  0  at  (0,  0)  and  a  maximum  of 
1  at  (1,0). 


y 


(c)  Vg  =  0  at  (0,  0) 
43.  (a)  Hint:Checkthat3L/3/,-  =  c;  -  &(x)fon  =  \,...,k 
and 

dL_  =  df_  _yj,dg[_ 
dxj       dxj      4^  dxj 

for  j  =  1, . . . ,  n. 

Section  4.4 

1 .  5x  -  ly  +  14  =  0 

3.  (a)  D(a,  b)  =  E?=i(y«  -  (a/x,  +  b)f 

(b)  Minimize  D  with  respect  to  a  and  b. 
5.  Hint:  Let  D(a,  b,  c)  =  IXiCv.  -  (ax2  +  bxt  +  c))2 . 
7.  (b)  There  is  a  single,  stable  equilibrium  point  at 

H.-i). 

9.  Single  equilibrium  point  at  (—  1 ,  |,  |).  There  are  no  stable 
equilibria. 


Answers  to  Selected  Exercises  589 


1 1 .  Produce  50,000  each  of  both  the  standard  and  executive 

models  and  100,000  deluxe  models. 
13.  Irrigate  only;  Purchase  3333.3  gal  of  water. 
1 5.  (a)  Invest  $120,000  for  capital  equipment  and  $240,000 
for  labor. 

(b)  Hint:  Note  that  L/K  =  2. 

True/False  Exercises  for  Chapter  4 

1 .  True 
3.  True 
5.  True 

7.  False.  (/  is  most  sensitive  to  changes  in  y.) 
9.  False 
1 1 .  True 

1 3.  False.  (Consider  the  function  f(x,  y)  =  x2  +  y2.) 
15.  True 

1 7.  False.  (The  point  is  not  a  critical  point  of  the  function.) 
19.  True 

21 .  False.  (The  critical  point  is  a  saddle  point.) 
23.  False.  (Extrema  may  also  occur  at  points  where  g  =  c  and 
Vg  =  0.) 

25.  False.  (You  will  have  to  solve  a  system  of  7  equations  in 

7  unknowns.) 
27.  True 

29.  False.  (The  equilibrium  points  are  the  critical  points  of  the 
potential  function.) 

Miscellaneous  Exercises  for  Chapter  4 

1 .  r0  =  2h0 

3.  Price  the  Mocha  at  $2.70  per  pound  and  the  Kona  at  $5 
per  pound. 

5.  Maximum  value  of  4  at  (1,  — V3,  0).  Minimum  value  of 

-4at(-l,  V3,  0). 
7.  (e) 


(c)  Maxima  at  Q=,  ^,oV  (- 


j_ 

V2' 


J= ,  0^ ;  minima 


at  (.7l<-71'0)'  (-7I'7I'0);  saddle  P°mts  at 
(0,0,  ±1) 

11.  l/(a2  +  a2  +  .--  +  a2) 

13.  Dimensions  are  4  (jc-direction)  by  2\/2  (y-direction)  by 

2  (z-direction). 
15.  1 

17.  a/3  by  6/3  by  c/3 
19.  3V5/8 

21 .  Hint:    Minimize    D2  =  (x  —  xo)2  +  (y  —  yo)2,  where 

(x,  y)  denotes  a  point  on  the  line  ax  +  by  =  d. 
23.  (a)  Hint:  Show  that  the  maximum  value  occurs  when 
x2  =  y2  =  z2  =  a2/3. 

(b)  Since  f(x,  y,  z)  =  x2y2z2  is  maximized  when  x2  = 
y2  =  z2  =  a2/3,  we  must  have  x2y2z2  <  (a2/3)3  = 
((*2  +  y2  +  z2)/3)3. 

(c)  Hint:  Since  x\,  Xz,  ■  •  ■ ,  x„  are  assumed  to  be  posi- 
tive, we  can  write  X;  =  yf  for ;  =  1 .  Maximize 

f(yi  ,yi,---,yn)  =  y\y\  ■■■yl  subject  to  y\  +  y\  + 


25.  (a)  A,i,  A2 


(a  +  c)±  y/(a  +  c)2  -  4(ac  -  b2) 


(b)  Rewrite  as  k\ ,  ^2 


Chapter  5 


(a  +  c)  ±  V '(a  -  c)2  +  4fo2 


9.  (a)(0,O,±l),(il,0))( 


j  l  rA 


(■ 


V2' 


Section  5.1 


40 

3 


1. 


3.  6(e  -  1) 


590       Answers  to  Selected  Exercises 


2e2  +  e  +  21n2 


7.  (a)  Volume  =      f^(x2  +  y2  +  2)  dy  dx  =  26 


9 
11 


(b)  Volume  =  /Q2  f^(x2  +  y2  +  2)  dy 
Jo  I-ii^x2  +  y4  sin^jc)dy  d.v  =  2  +  66/5^ 


The  iterated  integral  gives  the  volume  of  the  region 
bounded  by  the  graph  of  z  =  16  —  x2  —  y2,  the  xy-plane, 
and  the  planes  x  =  1,  x  =  3,  y  =  —2,  y  =  2.  The  value 
of  the  integral  is  . 
13.  The  iterated  integral  gives  the  volume  of  the  region 
bounded  by  the  graph  of  z  =  4  —  x2,  the  jcy-plane,  and 
the  planes  x  =  —2,  x  =  2,  y  =  0,  y  =  5.  The  value  of  the 
integral  is  . 

15.  The  iterated  integral  gives  the  volume  of  the  region 
bounded  by  the  graph  of  z  =  5  —  \y\,  the  xy-plane,  and 
the  planes  x  =  —I,  x  =  2,  y  =  —5,  y  =  5.  The  value  of 
the  integral  is  75. 


Section  5.2 

1.  0 


3.  (a)  Check  that  f\fn  "  x3dydx 


0. 


(b)  Hint:  The  region  D  is  symmetric  about  the  y-axis  and 
x3,  is  an  odd  function. 


5.  4 


2 

(4,2) 

1  - 

— i — i — i — i — 

12       3  4 

7  1^2 
'  ■  3 


(-1,-1)'-! 


9.   1  —  COS  16  y 
4 


11.  3n/2  y 


1- 

Md.o) , 

J  2 

-1- 

13.  0 


V  =  e* 

2- 

i 

D 

,  (1,  e) 

 1  

-1 

 1  1  

2       3  4 

-1  - 

\ 

y  =  -ex 

4 
3 

99 
20 


128 

5 


15. 

17. 
19. 
21. 

23.  Hint:  Write  / /s  cf  dA  and  ffRf  dA  as  limits  of  appro- 
priate Riemann  sums. 

25.  Hint:  Use  the  fact  that  |^"=i  a<  \  —  \a'\- 

27  i 

29.  Area  is  7rafo. 

31  1 

6 

18 


33 
35 
37 
39 


50 
3 


11,664 


16 

3 


Answers  to  Selected  Exercises  591 


41.  (a)  L  f(x,  y  )dy  =  2  regardless  of  whether  x  is  rational 
or  irrational. 

(b)  Value  is  2. 

(c)  2 

(d)  Converges  to  3. 

(e)  The  Riemann  sum  has  no  uniquely  determined  limit. 


(2,4) 


(c)  f*f£(2x  +  \)dxdy  =  4 


r4  r2-y/2 
10  JO 

y 


y  dx  dy 


16 

3 


4-" 

(0,4) 

3- 

2- 

V 

=  4-2x 

1- 

 1 — 

\(2,0) 

1 

2 

5.  fifi(x  +  y)dydx=% 
(3,9); 


1-  Jo  £/2  eX  dy  dx  +  fi  fl/2  e"  dy  dx  =  j  +  e>/2  -  e 

y 


\x  =  l 

1  - 

x  =  y  - 

As —  i 

^  1  1  

1  2 

9-         C  '  '  ydydx 


16 
3 


(0,2) 

=  -M-y2/^ 

=  V4  -  y 

/  1- 

(-2,0)  ( 

\(2,0) 

-2  -1 

1  2 

11  625 
II.  12 

1  ^  896 

io.  15 

15.  (1  -  sinl)/2 
17.  (e27  -  l)/6 

19.  (a)  If  your  computer  algebra  system  can  provide  a  simple 
answer,  it  should  be  4(1  —  cos  2). 

(b)  The  integral  with  respect  to  y  requires  two  applica- 
tions of  integration  by  parts. 

(c)  It  should  take  the  computer  only  a  fraction  of  a  second 
to  find  that  J*.1  f^y  y2  cos  (xy)  dx  dy  =  \  ( 1  —  cos  2), 
much  faster  than  the  evaluation  in  part  (a). 

21.  (a)  It  is  quite  possible  that  the  computer  will  be  unable  to 
make  the  evaluation, 
(b)  With  order  reversed,  the  computer  easily  finds 
J^fre—dydx  =  e-L 

Section  5.4 

1.  0 
3.  1 

5 
7 


1539 
16 
__5_ 
24 


9.  Volume  : 


11  12§ 
ii.  15 


ra     rVa2-.v2  r^Ja2-x2-y2 

LaL^?I_^-^  dzdydx 


13.  0 

15  i 

'  10 

17.  81V3tt/8 


592       Answers  to  Selected  Exercises 


19  15 
21  ?9 

3 

23.  V2tt/2 

25.  Jo      Jo  ~A  /(*>  y  •  z)  rfz  rfy 

=  Jo'  fo~X  fl%  f(*>  y.  z)dydzdx 
=  fo  fo~Z  11%  f(x>y>  z) dydxdz 

=  f-i  fo~}  If"  /(*•  y> z)dx dz dy 

=  fo  f-%=i  If'  f(x,y,z)dxdy  dz 


z 


X 


27  •  fo  fy  fo  f(x'  y '  z)  dzdxdy 

=  ^  fo  fz  f(x>  y.  z)  dydz  dx 
=  ^ fz  fz  f(x,y,z) dydxdz 

=  f02  fo  f2  f(x>  y>  z)dx dz dy 
=  fo2  f2  f2  f(x<  y> z> dx  dy dz 

29.  (a)  Bottom  surface  is  z  =  x2  +  3y2,  top  surface  is  z  = 
4  —  y2;  shadow  in  xy-plane  is  x2/4  +  y2  <  1,  y  >  0. 

00  fof2_i^j:i:^  +  y')dzdxdy 

(c)  /0734//_^J(^3  +  y3)^^^y 

(d)  fo  If*  ./^;:;<'S  I  Vs), A  rfy  dz  + 

(e)  /_22  /,r2/4  /./^V  +  y3)  ^  *  «fa  + 
X?2  ft+xVi  fo  4^(*3  +  y3)  Jy dzdx 

Section  5.5 


1 .  (a)  T(m,  v)  = 


'  3 

0  " 

u 

0 

-1 

V 

(b)  D  =  [0,3]  x  [-1,0] 
3.  D  is  the  parallelogram  with  vertices  (0,  0),  (11,  2),  (4,  3), 
(15,5). 


5.  T  takes  W*  onto  the  parallelepiped  with  vertices 
(0,  0,  0),  (3,  1,  5),  (-1,  -1,  3),  (0,  2,  -1),  (2,  0,  8), 
(3,3,4),  (-1,  1,2),  and  (2,2,  7). 

7.  (a)  D  =  {(x,  y,  z)  |  x2  +  y2  +  z2  <  1} 

(b)  D  =  {(x,  y,  z)  |  x2  +  y2  +  z2  <  1,  x,  y,  z  >  0} 

(c)  D  =  {(x,  y,  z)  |  1  <  x2  +  y2  +  z2  <  1,  x,  y,  z  >  0} 

9.  |02  /02  U5i;e"2  -\dudv  =  8(e4  -  l)/3 
11.  7(e-  1/e) 
13.  3tt 
15.  486tt/5 
17.  31n(^+  1) 
19.  (16  -  3tt)/12 
21.  na2/4  if  n  is  odd, 
7ta2/2  if  n  is  even. 
23.  2  +  JT/4 
25.  (jrsinl)/3 
27.  l(ln(V2+ 1)  + V2- 1) 
29.  7t(e4  -  e  +  1) 
31.  48tt 

33.  ^[a3  -(a2  -fo2)3/2] 

35.  2tt((1  -  a2)eal  +  (b2  -  \)ebl) 

37.  8505tt/32 

39.  4V2tt(5^-8)/3 

41.  656tt/5 

Section  5.6 

1 .  (a)  80  cases 

(b)  SI. 60 
3.  e2  -  2e  +  1 

5.  (a)  c,  where  c  is  the  constant  of  proportionality, 
(b)  {(x,y,z)  |  x2  +  y2  +  z2  =  1} 
30-3tt 

30-5tt 
9.  90  seconds 

1 1 .  If  the  plate  is  located  at  {(x,  y)  |  x2  +  y2  <  a2,  y  >  0}, 

then(x,y)  =  (0,  4a/3jr). 
13.  (x,y)=(f,^) 

1  5.  (x,  y)  =  ((7V3  +  8jt)/(3V3  +  4tt),  15/(Vl  +  An)\ 

17.  (jc,  50  =  (21/20,0) 
19.  (a)  (*,y,z)=(i,0,§) 

(b)  (x,y,z)=(g,0,  |) 
21.  (x,y,f)=(0,0,-l|) 
23.  3o/8 

25.  (a)  /v  =  /,,  =  If  =  ^ 

(b)  r,  =  r„  =     =  1/V5 

27.  (a)  /-  =  6561ir/4;  r-  =  (3-v/3)/(2V2) 
(b)  /,  =  8748^/35;  r,  =  (373)777 


Answers  to  Selected  Exercises  593 


29.  1496/135 

31.  116tt/3 

33.  V  =  -3GMm(b2 

Section  5.7 


■a2)/(2(fc3-a3)) 


1.  (a)  J2,3  =  -0.621375 
b)  Exact  value  is  —0.619. 

a)  T2,i  =  0.19825 

b)  Exact  value  is  0.197402. 

a)  72,3  =  0.336123 

b)  Exact  value  is  0.33 1642. 

7.  (a)  52,2  =  0.00010125 

b)  Exact  value  is  0.0000972. 

a)  52,2  =  0.331871 

b)  Exact  value  is  0.331642. 

11.  (a)  S2,2  =  0.414325 

b)  Exact  value  is  0.414214. 

.  (a)  io1  fo  y/4x2  +  36y2  +ldydx 
b)  J4,4  =  3.52366 
1 5.  52,2  is  more  accurate  than  T4  4. 

17.  (a)  n  must  be  at  least  19. 
b)  n  must  be  at  least  1 . 

19.  (a)  S2,i  =  -0.00390625 

b)  Exact  value  is  —  1  /256. 

c)  The  answers  are  the  same  because  £2,2  =  0  by  Theo- 
rem 7.4. 

21.  7-3,3  =  0.190978 

23.  73,3  =  0.412888 

25.  7-3,3  =  0.724061 

True/False  Exercises  for  Chapter  5 

1 .  False.  (Not  all  rectangles  must  have  sides  parallel  to  the 
coordinate  axes.) 

3.  True 

5.  False 

7.  False 

9.  True 

1 1 .  True 

1 3.  False.  (The  value  of  the  integral  is  3.) 

15.  True 

1 7.  True.  (The  inner  integral  is  zero  because  of  symmetry.) 

19.  True 

21 .  False.  (The  integrals  are  opposites  of  one  another.) 

23.  False.  (A  factor  of  r  should  appear  in  the  integrand.) 


25.  False.  (A  factor  of  p  is  missing  in  the  integrand.) 
27.  True 
29.  True 

Miscellaneous  Exercises  for  Chapter  5 

1.  72;r 


(a)  f0  3  L%i%s  fi/+y2  3  dz  dy  dx 


-3  io        J2r2+v2  ^  "z    "y are  two 


and 
possibilities, 
(b)  8lV3  7r/4 

-1    rVT^  p/9 


dzdy  dx 


o»  ,/;r  ,c 

1/3  foSC<P  P1  sin(pdp  dq>  d6.  Value  is 
16V2N 


flit  fim  1  1/3  p~i    2    ■        j    j     jr\  1 

J0  p  smtp  dp  d(p  d0  + 

rliz  rn/2 
JO  Jsi 


2jt  9 


7. 


fo  fo/4  io"SeCS  rdr  de  dz 


9.  (sinl  +  sin2)/6 


11.  (a)  r  f^E^Ldydx 


(b)  E* 


{(x,  y)\x2  +  y2 


(c)  Area  =  f^  fQl  abr  dr  d6 


13.  Area  is  it. 

15.  -fln4 

17.  (In2)(tan-14 


-) 
4  I 


y2  rV^ 


dw  dz  dy  dx 


19.  ^f-Jl^f^ 

(b)  7t2a4/2 

(c)  Five-dimensional  ball  has  volume  8jr2a5/15;  six- 
dimensional  ball  has  volume  jr3a6/6.  The  pattern  is 
not  very  clear  from  this  information. 

25.  Within  a  disk  of  radius  V5a/3  about  the  center,  where  a 
is  the  radius  of  the  hemisphere. 

27.  (a)  4(VT^7  -  v^XvT3*  -  V8) 
(b)  4 

29.  Integral  does  not  converge. 
31.  -4ir/9 

33.  Converges  when  p  >  1  and  q  >  1;  value  is 
1 


(p-\)(q-\) 

35.  Integral  does  not  converge. 

37.  (a)  Hint:  Break  up  the  integral  into  a  sum  of  integrals 
from  0  to  1  and  from  1  to  00.  Show  convergence  of 
the  improper  integral  from  1  to  00  by  comparing  it  to 

f^°  e~x  dx. 


594       Answers  to  Selected  Exercises 


(b)  Hint:  Begin  with  JX.,e  x   y  d  A  and  use  laws  of  ex- 
ponents. 

(c)  7r(l-e-fl2) 

(d)  7T 

(e)  ^/n 
39-  (b)  4 
41.  C  =  a6/4 
43.  (a)  C  =  ^ 

(b)  | 

45.  1  -e-'/t  0.2212 


Chapter  6 

Section  6.1 

1.  (a)  50       (b)  4 

3.  (35V35  -  17\/l7)/27 

5.  5vT7 

7.  (53/2  +  87)/12 

9.  2  11.  |  13.  (3e3  -5)/3 

15.  4tt  +  16;r2/3 

17.  3tv  19.  1  -  e"4;r 

21.  (a)  /xF.ds=i,/yF.ds  =  -i 

(b)  y(r)  =  x(l  —  2f).  Thus,  the  images  of  x  and  y  are  the 
same,  although  y  traces  the  image  in  the  opposite  di- 
rection to  that  of  x.  For  these  reasons,  the  results  of 
part  (a)  could  have  been  anticipated. 

23.  0 

25.  -M  27.  9  29.  0 

31.  -li 

33.  Hint:  Use  formula  (3). 
35.  (a)  250(k/T3  ft-lb 

(b)  7500  ft-lb 
43.  (a)  7.65625       (b)  7.5 

Section  6.2 

1 .  §dDMdx  +  Ndy  =  f fD(Nx  -  My)dA  =  8tt 
3.  <fSD  M  d.x  +  Ndy  =  f  fD(Nx  -  My)dA  =  -4 
5.  fdD  Mdx  +  N  dy  =  f fD(Nx  -  My)dA  =  -  Ujljz 
7.  (a)  0 

(b)  You  will  need  to  compute  four  separate  line 
integrals — one  for  each  edge  of  the  square. 

9.  -2 

1 1 .  Hint:  Calculate  \  fc  —y  dx  +  x  dy,  where  C  is  the  bound- 
ary of  R,  oriented  counterclockwise. 


13.  -45 

15.  (a)  y 


\  1 

-1  / 

/  1 

CO  TJ 

1 9.  Hint:  Parametrize  each  line  segment  of  the  perimeter. 
21.  -12jt 

23.  (a)  Hint:  Show  V  •  F  =  0. 
25.  Directly  apply  Green's  theorem. 

27.  The  line  integral  is  ±3  times  the  area  of  the  rectangle, 
where  the  sign  depends  on  the  orientation  of  the  boundary. 

29.  Hint:  uVu  =  (u  dv/dx,  u  dv/dy).  Then  use  Green's 
theorem. 

31.  Hint:  Begin  with  (b    — ds  and  then  use  Green's 
Jsd  dn 

theorem. 

Section  6.3 

1.  (a)  f 

(b)  ^ 

(c)  No.  Line  integrals  are  not  path-independent. 
3.  Not  conservative 

5.  Not  conservative 

7.  Conservative;  f(x,  y)  =  xe~y  +  cosjfy 
9.  Conservative;  f(x,  y)  =  3x2y2  —  jc3  +  |y3 
1 1 .  Conservative;  f(x,  y,  z)  =  2x2yz^  —  x2y  +  y2z 
1 3.  Conservative;  f(x,  y,  z)  =  x2  +  xy  +  sinyz 
1  5.  Conservative;  f(x,  y,  z)  =  ex  sin  y  +  z3  +  2z 
17.  Not  conservative 

21 .  N(x,  y)  =  je2x  +  x3ey  +  u(y),  where  u  is  any  function 

of  y  of  class  C2. 
23.  N(x,  y,  z)  =  \x4  +  z2  +  h(y),  h  of  class  C1. 

25.  (a)  Check  that  V  x  F  =  0.  Since  F  is  of  class  C1  on  all 
of  R3,  by  Theorem  3.5,  F  is  conservative.  A  scalar 
potential  is  f{x,  y,  z)  =  ^x3  +  sin  y  sin z. 
(b)  |  +  sine  sine2  —  (sin  l)2 
27.  0 
29.  6 
31.  -480 

33.  (a)  Conservative  on  {(x,  y)  \  y  >  0}  and  on 
{(x,  y)\y  <  0} 


Answers  to  Selected  Exercises  595 


(b)  f(x,y)  =  (x2y2+x2  +  l)/(2y2) 

(c)  1 

35.  (a)  F  is  conservative.       (b)  —2 

37.  Work  =GMm(—  —  ) 

Vllxill  ||x0||/ 

True/False  Exercises  for  Chapter  6 

1 .  True 

3.  False.  (It's  negative.) 

5.  False.  (The  integral  is  0.) 

7.  False.  (There  is  equality  only  up  to  sign.) 

9.  True 
1 1 .  True 
13.  True 

1 5.  False.  (The  line  integral  could  be  ±  fc  ||F||  ds,  depending 
on  whether  F  points  in  the  same  or  the  opposite  direction 
as  C.) 

1 7.  False.  (Let  F  =  yi  —  xj  and  consider  Green's  theorem.) 
1 9.  False.  (Under  appropriate  conditions,  the  integral  is 

f{B)  -  f(A).) 
21.  True 

23 .  False.  (For  the  vector  field  to  be  conservative,  the  line  inte- 
gral must  be  zero  for  all  closed  curves,  not  just  a  particular 
one.) 

25.  False.  (The  vector  field  (ex  cos  y  sin z,  ex  sin  y  sin z, 

ex  cos  y  cos  z)  is  not  conservative.) 
27.  False.  (The  domain  is  not  simply-connected.) 
29.  False.  (/  is  only  defined  up  to  at  most  a  constant.) 

Miscellaneous  Exercises  for  Chapter  6 

1.  Break  up  C  into  n  segments,  each  of  length  A.sj.  By 
continuity,  /  will  be  nearly  constant  on  each  segment, 

so  [/]avg  ~  ELi  f(ck)&Sk/ ELi  Ask-  (Here  c*  is  any 
point  in  the  kth  segment.)  The  formula  follows  after  taking 
limits  as  all  As*  — >  0. 

3.  la/n 

5.  2 

7.  x  =  y  =  (*-*ri"-2f)a 

y        \    A(n-2sfl)  ) 

9.  (a)  (b)  2 

11     j   _  7769^17-1  _    /  7769Vf7-T 

1  ~~        840      '  r*  -  V  1190^17-70 

13.  /-  =27711,        rz  = 

15.  (a)  Xf  g(f{e)  cos  6,  f(6)  sinfl V(/(0))2  +  {f'(6)f  d6 

(b)  yi0[(e1811  -  l)/9+  12(1  -e12ir)/37] 

17.  6tt 

19.  K  =  2n 


21 .  |  —  cos  1  —  sin  1 

23.  Hint:  Use  the  formula  Area  =  j  J3D  —y  dx  +  x  dy. 
25.  Hint:  x  =  ffDx  d  A/area  of  D, 

y  =  ffDy  dA /area  of  D.  Now  apply  Green's  theorem  to 

<fdD  x2  dy  and  §aD  xy  dy. 

27.  x  =  y  =  -±4 

29.  Hint:  Use  the  result  of  Exercise  28  twice. 
35.  (a)  Both  the  line  integral  and  the  double  integral  are  zero, 
(b)  No.  The  double  integral  is  not  defined  properly  over 
the  disk  because  F  is  undefined  at  the  origin. 

W  7/fl[^(^)-i(^)]^  =  o  =  /cF-^  + 

fc  F  •  ds.  Since  fc  F  •  ds  =  —  2n,  the  result  follows. 
37.  (a)  0  " 

(b)  Apply  the  divergence  theorem. 
39.  /xF- ds  =  |X-VV  •  ds  =  -V(B)+  V(A).    Now  use 

Exercise  38. 

Chapter  7 

Section  7.1 

1.  (a)  -i-4j  +  2k 

(b)  x  +  4y  -  2z  =  5 
3.  y  +  4z  =  4 


y 


(b)  Yes 

(c)  4x  -  2y  -  z  =  3 

7.  (a)  Smooth  except  at  (0,  0,0).  Tangent  plane  at  (1,V3,  4) 
has  equation  2x  +  2^/3y  —  z  =  4. 

(b)  S  is  a  paraboloid. 

(c)  z  =  x2  +  y2 

(d)  Yes  it  does.  At  (0,  0,  0),  the  tangent  plane  has  equation 
z  =  0. 

9.  Hint:  Consider  x2  +  y2. 
1 1 .  (a)-(c)  All  versions  give  the  equation 

-x  +  y  +  \/2z  =  1. 
13.  X(s,  t)  =  (2cos  s,  r,  2 sins),  0  <  s  <  2jt,  -1  <  t  <  3 

15.  Xi(s,  t)  =  (s,  t,  Vs2  +  t2+  lY 

X2(s,  f)  =  ^,  f,  -V.?2  +  f2  +  l),  (i,    G  R2 

17.  (a)  y2z  =  t2s2=x2 


596       Answers  to  Selected  Exercises 


x 


(d)  Points  on  the  positive  z-axis 

(e)  -Ax  +  Sy  +  z  =  4 

(f)  (0,  0,  1)  =  X(±l,  0);  tangent  planes  have  equations 
x  ±  y  =  0. 

19.  (a)  At  the  point  (a,  g(a,  c),  c),  the  tangent  plane  has  equa- 
tion gx(a,  c)(x  -a)-(y-  g(a,  c))  +  gz(a,  c)(z  - 
c)  =  0. 

(b)  At  the  point  (h(b,  c),  b,  c),  the  tangent  plane  has  equa- 
tion -(x  -  h(b,  c))  +  hy(b,  c)(y  -  b)  +  hz(b,  c){z  - 
c)  =  0. 

21.  {x  e  R3  |  x  =  (1,  0,  \)  +  (s  -  1)(1,2,  0) + 

0  +  1)(0,  1,  -2)}  or  x=s,y  =  2s  +  t-\,z  =  -It  - 
1.  To  verify  consistency  with  Exercise  5(b),  check  that 
the  points  given  by  the  parametric  equations  all  lie  in  the 
plane  determined  by  the  equation  in  Exercise  5(b). 

23.  V6tt 

25.  4tWo2  -  b2 

27.  (653^2  -  173/2)tt/24 

29.  yTTa -areaofD 

31.  16a2 

Section  7.2 

1.  26V3/3 
3.  1 

r  640 
O.  3 

7.  (a)  4jra4        (b)  47za4/3 

9.  (a)  Parametrize  the  cylinder  as  x  =  2cosf,  y  =  2sinf, 
z  =  s  with  —2  <  s  <  2,  0  <  /  <  lit .  The  integral 
evaluates  to  —  647r. 
(b)  The  integral  is  —4  •  (surface  area  of  S)  =  —64n. 
11.  0 

13.  297tt/2 
15.  36tt 
17.  0 

19.  27rfl3/3 
21.  —Tta2 


(b)  ni  =  (x/(V2z),  y/(V2z),  -1/V2),  n2  =  (x,  y,  0) 

(c)  -1456jt/3 
29.  (a) 


z 


(b)  (a  cosi,  a  sins,  0) 

(c)  N(s,  0)  =  (a  cos  s  (2  cos(s/2)  +  sin(s /2)), 

a  sins  (2 cos(s/2)  +  8^(^/2)), 

a(2  sin(s/2)  —  cos(.?/2))). 

From  this,  we  see  that  N(0,  0)  =  (2a,  0,  —a),  while 
N(2tt,  0)  =  (-2a,  0,  a).  Therefore,  the  Klein  bottle 
cannot  be  orientable,  since  the  normal  vector  along 
the  s-coordinate  curve  at  t  =  0  changes  direction. 

Section  7.3 

1.0  3.  0 

5.  0  7.  0 

9.  4jr(b2  -  a2) 
11.  45tt 

13.  (a)  Hint:  Use  the  double  angle  formula. 

(b)  -f 
15.  1/2 
17.  625tt/8 
25.  (a)  na 

(b)  it  a 

(c)  The  answers  in  parts  (a)  and  (b)  are  the  same;  the 
three  flat  quarter-circles  that  are  part  of  3D  do  not 
contribute  anything  to  jjsVf-ndS. 


Answers  to  Selected  Exercises  597 


29.  (a)  §sF-dS 


«  Fz(r,  9,z  +  Az/2)r  A9  Ar 

-  Fz(r,  6,z-  Az/2)r  AO  Ar 

+  Fr(r  +  Ar/2,  9,  z)(r  +  Ar/2)A9  Az 

-  Fr(r  -  Ar/2,  9,  z)(r  -  Ar/2)A9  Az 
+  F9(r,  9  +  A6/2,  z)Ar  Az 

-  Fe(r,  9  -  A6/2,  z)Ar  Az 

Section  7.4 

1 .  Hint:  Use  Gauss's  theorem  and  the  product  rule  for 
V-(/Vg). 

3.  (a)  Hint:  Use  Green's  first  formula  with  f  =  g. 

(b)  Hint:  Use  part  (a)  and  the  fact  that  V/  •  V/  = 
IIV/H2. 

7 .  Hint:  Use  the  argument  in  Exercise  6  and  the  product  rule 

forV-(A;Vr). 
9.  (a)  Hint:  Use  Gauss's  theorem  and  Exercise  8. 

(b)  Heat  flows  into  D  from  the  inner  sphere  and  out 
through  the  outer  sphere  at  the  same  rate. 
1 1 .  Hint:  Use  Ampere's  and  Gauss's  laws. 
15.  (b)  In  each  case  k  =  [jlq€q. 

1 9.  Hint:  Apply  Gauss's  theorem  to  P  =  E  x  B,  then  Fara- 
day's and  Ampere's  laws. 

21.  Hint:  Apply  the  arguments  used  in  Exercise  20  to  each 
component  integral  of  B. 


(a)  (Vz2+  1  cos  (Wz2  +  1  sin6»,  z) 

(b)  (a*Js2  +  1  cos?,  b\l s1  +  1  sin?,  cs) 
(a)  (a  sirup  cos9,  bsincp  sinO,  c coscp) 


(bU2Vo* 


b2c2  sin4  <p  cos2 

+a2b2  cos2  cp  sin2  (p 


+  a2c2  sin4  cp  sin2  ( 


dtp  dG 


(a)  (s  cos  t,  f(s),  s  sinf) 

(b)  t 


n 

10 


11. 

13.  (a/2,  a/2,  a/2) 
15.  (0,0,  a/3) 
17.  (a)  15>/57r/2 

(b)  V572 

(c)  Iz  =  62V5jrk/5,  rz  =  793/35 
19.  (a)  Ix  =  Iy  =  2jrab8(3a2  +  2b2)/3 

(b)  r,  =  ry  =  V(3fl2  +  2b2)/6 
23.  Hint:  Use  Stokes's  theorem. 
25.  Use  Exercise  24. 

31 .  Hint:  Show  V  •  (x/||x||3)  =  0  where  defined. 
35.  V  •  F     0,  so  there  is  no  vector  potential. 
39.  Hint:  Calculate  V  x  (E  +  dA/dt). 


Chapter  8 


True/False  Exercises  for  Chapter  7 

1 .  True 

3.  True.  (Let  u  =  s3  and  v  =  tanf.) 

5.  False.  (The  limits  of  integration  are  not  correct.) 

7.  False.  (The  value  of  the  integral  is  24.) 

9.  True 

1 1 .  False.  (The  integral  has  value  32jt.) 

1 3.  False.  (The  value  is  0.) 

1 5.  False.  (The  surface  must  be  connected.) 

17.  True.  (The  result  follows  from  Stokes's  theorem.) 

19.  True.  (Use  Gauss's  theorem.) 

21 .  False.  (Gauss's  theorem  implies  that  the  integral  is  at  most 

twice  the  surface  area.) 
23.  True 

25.  False.  (Should  be  the  flux  of  the  curl  of  F.) 
27.  True.  (Apply  Green's  first  formula.) 
29.  False.  (/  is  determined  up  to  addition  of  a  harmonic 
function.) 

Miscellaneous  Exercises  for  Chapter  7 

1.(a)  C       (b)  E       (c)  A 
(d)  D        (e)  F        (f)  B 


Section  8.1 

1.-2  3.  6  5.  182 

7.  -370 

9.  —6a\b^  +  6a^b\  +  02^4  —  04^2 

11.  14cosx  -  7sinz  +  lly2  +  33 

13.  6  (ev  cos  y  +  (y2  +  2)e2z) 

1 5.  2xy  dx  A  dy  A  dz 

17.  3^3(^:3  —  X4)  dx\  A  dx2  A  dx-$  A  dx4 

1 9.  (x\eXAXs  —  X1JC2X3  cos X5)  dx\  A  dxi  A  dx^  A  dx$  A  dx^ 

Section  8.2 

1.  Hint:  Show  linear  independence  of  T9l,  Tg2,  Tg3  by 
solving  the  vector  equation  CiT^,  +  C2T#2  +  c^TS]  =  0 
for  c\,  C2,  C3. 

3.  X:  [0,  2tt)  x  [0,  2jt)  x  [0,  2jt)  x  [1,  2]  x  [1,2]  R6, 
X(6»i,  62,  6*3,  l2,  h)  =  (xu  y\,  x2,  V2,  xi,  y3),  where  x\  = 
3  cos 9\,  yi=3sin#i,  X2  =  3  cos#i  +  h  cos 6*2,  y2  = 
3  sin 61  + 12  sin 62,    X3  =  3  cos#i  +  h  cos  02  +  h  cos03, 
y3  =  3  sin  9\  +  1%  sin  #2  +  ^3  sin  #3 

7.  -2jt 

9.  4jt2 

11.  (a)  Sx(u)(T„1,T„2,T„3)  =  m1  >  0    for    0  <  u\  <  JJ 
which  is  where  the  parametrization  is  smooth. 


598       Answers  to  Selected  Exercises 


(b)  Parametrize  dM  in  two  pieces  as 
Yi:[0,V5]  x  [0,  2tt)  R3, 
Yi(ii,  S2)  =  (si  cosi2,  *i  sinj2,  $\  —  6)  and 


(c)  V 


Y2:[0,  V5]  x  [0,  lit)  R3, 

*2)  =  (*1  COS  52, 

(2«i  cos  S2 ,  2s  1  sin  ^2, 


Y2CS1,  52)  =  Csi  COS52,  *i  sinj2,  4  —  ij). 


1) 


V2 


lAs\  +  1 
(2*i  cos 52,  25i  sin 52,  1) 


4j?  +  1 


13.  32?r 

15  ^ 
u.  2 


Section  8.3 

1 .  dco  =  exyz(yz  dx  +  xzdy  +  xy  dz) 

3.  dco  =  —ydx  A  dy 

5.  dco  =  (x  +  2yz)<ix  A  dy  A  dz 

7.  d&>  =  2(X]  —  X2  +  X3  —  •  ■  •  +  (—  l)"+1x„) rfxi  A  JX2  A 
•  •  •  A  dx„ 


9.  F(x,  z)  =  xz  +  Cz  +  D\,      G(x,y)  =  xy  +  Cy  +  D2, 
where  C,  D\,  D2  are  arbitrary  constants. 


True/False  Exercises  for  Chapter  8 

1 .  True  3 .  True 

5.  True  7.  True 

9.  True 

11.  False.  (X(l,  1, -1)  =  X(l,  1,  1),  so  X  is  not  one-one 
on  D.) 

1 3.  False.  (The  agreement  is  only  up  to  sign.) 
1 5.  False.  (This  is  only  true  if  n  is  even.) 
17.  True 

19.  True.  {dco  would  be  an  (n  +  l)-form,  and  there  are  no 
nonzero  ones  on  R".) 

Miscellaneous  Exercises  for  Chapter  8 

5.  0 

7.  (a)  V  x  (V/)  =  0 

(b)  V  •  (V  x  F)  =  0 
9.  Hint:  Consider  fM  d(w  Arj). 


Index 


n,  xv 

U,  xv 

V  (del  operator),  227-228 
C,  xv 

C,  XV 

A 

acceleration,  190 

normal  component  of,  2 1 5 

tangential  component  of,  215 
accumulation  point,  103 
addition  of  vectors,  see  vector(s), 

addition,  2 
adjoint  matrix,  61 
Ampere's  law,  516-517,  518 

static  case,  516 

time-varying  case,  517,  518 
angle  between  vectors,  2 1 ,  49 
angular  momentum,  243 
angular  speed,  34 
angular  velocity  vector,  221 
antiderivative,  xvii 
Archimedes'  principle,  506 
arclength  function,  205,  217 
arclength  parameter,  205-207 
area 

zero,  319 
area  element 

Cartesian,  364,  370 

general,  370 

polar,  364,  370 
area  of  a  surface,  464^167 
arithmetic  mean,  308 
arithmetic-geometric  inequality,  308 
astroid,  219 

average  value,  373-377 

of  a  function  on  a  surface,  525 

B 

ball 

closed,  101 

open,  101 
basis,  see  vector(s),  standard  basis,  9 
Bernoulli,  Johann,  194 
Bezier  curve,  239 
biharmonic  function,  187 
binormal  spherical  image,  242 
binormal  vector,  see  vector(s),  binormal, 

212 
BMI,  155 

body  mass  index,  155 
boundary 

of  a  manifold,  541 
boundary  of  a  set,  102 
Brahe,  Tycho,  193 
budget  hyperplane,  56 
buoyant  force,  506 


c 

Cauchy-Riemann  equations,  242 
Cauchy-Schwarz  inequality,  50,  293 
Cavalieri's  principle,  310 
center  of  mass 

continuous 
in  R,  380 
in  R2,  380 
in  R3,  382 

continuous,  379-383 

discrete,  377-379 

of  a  wire,  451 
centroid,  383 
chain  rule,  142-153 

in  one  variable,  143-144 

in  several  variables,  144-153 
change  of  variables 

in  double  integrals,  357,  370 

in  triple  integrals,  365,  371 
charge  density,  514 
charge  distribution,  514 
C°°,  138 

circular  helix,  190 
circulation,  413,  476 
circulation  density,  498 
C*  138 

Clairaut,  Alexis,  137 

class  C°°,  138 

class  C*,  138 

closed  box,  337 

closed  rectangle,  310,  314 

closed  set,  102 

clothoid,  241 

Cobb-Douglas  production  function,  302 

codomain  of  a  function,  82 

cofactor,  57 

column  vector,  52 

commodity  bundle,  56 

compact  set,  271 

component  functions,  85 

conduction  current  density,  517 

conductivity,  522 

connected,  441 

conservative  vector  field,  297,  441 
constraint,  279 
constraint  equation,  278 
continuity 

piecewise,  318 
continuity  equation,  516,  520 

for  current  densities,  520 

in  fluid  dynamics,  520 
contour  curves,  87 

definition,  88 
coordinate  axes 

in  R2,  xv 

in  R3,  xv 
coordinate  curves 


of  a  manifold,  538 
coordinate  transformations,  349-355 
coordinates 
Cartesian,  62 
conversions 
between  Cartesian  and  cylindrical, 
65 

between  Cartesian  and  polar,  64 
between  Cartesian  and  spherical, 
68 

between  cylindrical  and  spherical, 
68 

cylindrical,  65 

hyperspherical,  72-73 

on  R2,  xv 

on  R3,  xv 

on  the  real  line,  xv 

polar,  62 

rectangular,  see  coordinates, 
Cartesian,  62 

spherical,  66 
Copernicus,  Nicholas,  193 
critical  point,  182,  265 

constrained,  279 

degenerate,  268,  288 
cross  product,  27-36 

and  determinant,  3 1 

applications,  32-35 

definition,  28 

inR",  61 

properties,  29 
curl,  227,229-231,497 

in  cylindrical  coordinates,  233,  508 

in  spherical  coordinates,  235,  509 
current  density,  514 

conduction,  517 

displacement,  517 
curvature,  209-210,  216,  217 

total,  452 

radius  of,  240 
curve,  419 

convex,  452 

Bezier,  239 

closed,  419 

simple,  419 
cusp,  80 
cycloid 

curtate,  17 

prolate,  18 

D 

Darboux  formulas,  221 
Darboux  rotation  vector,  22 1 
definite  integral,  see  integral,  definite, 
xvi 

degree  of  a  polynomial,  141 


600  Index 


degree  of  a  term,  141 
del  operator  (V),  227-228 
derivative,  xv,  124,  127-128 
directional,  160,  161 
maximization  of,  163 
minimization  of,  163 
of  a  vector  field,  236 
exterior,  see  exterior  derivative,  553 
linearity  of,  134 

normal,  see  normal  derivative,  453, 

507,  520 
partial  derivative 

mixed,  136 

of  higher  order,  136-138 
partial  derivative  of  f  with  respect  to 
xt,  117 
determinant,  30,  56 

properties,  60 
differentiability 

and  continuity,  122,  127 
definition,  121 

in  general,  126 
of  a  function  of  three  or  more 

variables,  123-127 
of  a  function  of  two  variables, 
120-123 
differential  geometry,  202 
differential  operators,  152 
differentiating  under  the  integral  sign, 
405 

direction  cosines,  26 

directional  derivative,  see  derivative, 
directional,  160 

displacement  current  density,  517 

displacement  vector,  5 

distance 
between  parallel  planes,  46 
between  point  and  line,  45^16,  77 
between  point  and  plane,  77 
between  skew  lines,  46-47 

divergence,  228-229,  496 

in  cylindrical  coordinates,  233, 
507 

in  spherical  coordinates,  235,  508 
divergence  theorem,  see  Gauss's 
theorem,  493 

in  the  plane,  432 
domain  of  a  function,  82 
dot  product,  18-25,  36 

definition,  19 

properties,  19 
double  integral,  316 

in  polar  coordinates,  360-362 

E 

eigenvalue,  309 
eigenvector,  309 
electric  field 

of  a  continuous  charge  distribution, 
514 

of  a  single  point  charge,  512 


electromotive  force,  517 
elementary  region 

in  space,  340-343 
definition,  340 

in  the  plane,  321-327 
definition,  321 
ellipsoid,  94 

parametrized,  468,  525 
elliptic  cone,  94 
elliptic  paraboloid,  94 
endowment  vector,  56 
energy 

kinetic,  297 

potential,  297 
epicycles,  193 
epicycloid,  80,  437 
epitrochoid,  80 

equation  of  continuity,  see  continuity 

equation,  520 
equation  of  first  variation,  227 
equilibrium  point,  298 

stable,  298 

unstable,  298 
equipotential  line,  224 
equipotential  set,  224 
equipotential  surface,  224 
Euler's  formula,  188 
evolute,  240 

extension  (/ext),  322,  343 
exterior  derivative,  553 

of  a  £-form,  553 

of  a  0-form,  553 
exterior  product,  534-535 

properties  of,  535 
extrema 

absolute,  264 

constrained  local,  288 

global,  264,  270-274 

local,  263 

second  derivative  test,  268 
extreme  value  theorem,  27 1 

F 

Faraday's  law,  517-518 
Fenchel's  theorem,  452 
field 

scalar,  see  scalar  field,  222 

vector,  see  vector  field,  221 
first  variation 

equation  of,  see  equation  of  first 
variation,  227 
flow  line,  225 
flow  of  a  vector  field,  227 
flux,  432,  476 
flux  density,  498 
force 

buoyant,  506 
Frenet-Serret  formulas,  217-219 
Fresnel  integrals,  241 
frustum,  526 

Fubini's  theorem,  314,  319,  339 
functions,  82-94 


average  value,  374 

along  a  curve,  45 1 
codomain,  82 

component,  see  component  functions, 
85 

continuous,  109 

algebraic  properties,  1 1 1 
domain,  82 

extension  of  (/ext),  322,  343 
homogeneous,  187,  309 
injective,  83 
linear,  55 

mean  value  of,  374 

of  more  than  one  variable,  84-94 

graphing,  87-94 

quadric  surfaces,  see  quadric 
surfaces,  93 
one-one,  see  functions,  injective,  83 
onto,  see  functions,  surjective,  83 
partial,  1 16 
polynomial,  106 
potential,  223,  441 
range,  83 
scalar-valued,  84 
smooth,  138 
surjective,  83 
fundamental  theorem  of  calculus, 

xvii 

G 

gauge  freedom,  528 
Gauss's  law,  512-514,  518 
differential  form,  514 
integral  form,  514 
Gauss's  theorem,  433,  493^196 
implied  by  generalized  Stokes's 

theorem,  558-559 
proof  of,  503-505 
generalized  Stokes's  theorem,  see 
Stokes's  theorem,  generalized, 
553 

geometric  mean,  308 

global  extrema,  264,  270-274 

on  compact  regions,  270-274 
gradient,  124,  158,  227-228 

in  cylindrical  coordinates,  233 

in  spherical  coordinates,  235 
gradient  field,  223 
gravitational  potential,  388 
Green's  first  formula,  510 
Green's  first  identity,  453 
Green's  formulas,  510 
Green's  second  formula,  510 
Green's  second  identity,  453 
Green's  theorem,  429 

implied  by  generalized  Stokes's 
theorem,  557 

proof  of,  433-436 

vector  reformulation  of,  43 1 
Green's  third  formula,  510 
gyration,  radius  of,  385,  451,  526 


Index  601 


H 

harmonic  function,  142,  243,  453,  511, 
520 

head-to-tail  addition,  4 
heat  equation,  142,  520,  521 

uniqueness  of  solutions  to,  521 
heat  flux  density,  489,  520 
helicoid,  142,  468 
helix,  85,  190 
Hessian,  255 
Hessian  criterion 

for  constrained  extrema,  286-290 

for  extrema,  265-270 
Hilbert  matrix,  79 
homogeneous  function,  187,  309 
hyperbolic  paraboloid,  90,  94 
hyperboloid 

parametrized,  524 
hyperboloid  of  one  sheet,  94 
hyperboloid  of  two  sheets,  94-95 
hyperplane 

budget,  56 
hypersphere,  73,  167,  293,  308 
hyperspherical  coordinates,  72-73 
hypersurface,  73,  124 
hypocycloid,  80 
hypotrochoid,  80 

ideal  gas  law,  96,  158 

identity  matrix,  59,  177 

implicit  function  theorem,  168-170 

general  case,  170 
improper  integral,  405^406 
incremental  change,  249 
injectivity,  83 

inner  product,  see  dot  product,  19 

inR",49 
integrability,  316,  339 
integral 

definite,  xvi 

double,  316 
improper,  405 

in  polar  coordinates,  360-362 
improper,  405-406 
iterated,  3 1 1 
line 

scalar,  see  line  integral,  scalar,  409 

vector,  see  line  integral,  vector,  411 
of  a  differential  form,  536-542 
of  a  A>form  over  a  fc-manifold,  541 
of  a  2-form,  537 
properties  of 

linearity,  320 

monotonicity,  320 
triple,  338 

improper,  405,  522 

in  cylindrical  coordinates,  366-368 

in  spherical  coordinates,  368-370 
integration  by  parts  for  differential 

forms,  562 


intersection  of  sets,  xv 
inverse  function  theorem,  172 
inverse  matrix,  61 
involute,  15,  240 
isobars,  224 
isolated  point,  103 
isoquants,  302 
isotherms,  224 
iterated  integral,  311 

J 

Jacobi  identity,  39,  78 
Jacobian,  172,  356,  365 
judo,  6-7 

K 

A'-form 

basic,  533 

general,  533 
Kepler,  Johannes,  193 
kinetic  energy,  297 
Klein  bottle,  489 

L 

Lagrange  multipliers,  280-290 
lamina,  380 

Laplace's  equation,  141,  453 
Laplacian  operator,  157,  187,  236 

inversion  formula  for,  511 
law  of  conservation  of  energy,  454 
least  squares  approximation, 

293-297 
Leibniz's  rule,  405 
length,  204 
level  curves,  87 

definition,  88 
limit(s),  97-111 

algebraic  properties,  106 

geometric  interpretation,  103-104 

intuitive  definition,  98 

rigorous  definition,  99-100 

uniqueness,  106 
line 

parametric  equations  of, 

see  parametric  equations,  of  a 
line,  13 
line  integral 

differential  form,  415 

numerical  approximation  of,  421^126 

path-independent,  439,  440 

scalar,  409 

vector,  411 
line  segment,  76 
linear  combination,  543 
linear  independence,  284,  539 
linear  mapping,  see  mapping,  linear,  55 
linear  regression,  293 
linear  span,  543 
lines 

skew,  46 
Lorentz  force,  428 


M 

magnetic  field 

of  a  moving  charge  distribution,  515 

of  a  moving  point  charge,  514 
magnetic  monopoles,  516,  518 
manifold,  202,  538 

underlying,  538 

boundary  of  a,  541 

general,  551 

smooth,  539 
mapping,  linear,  55 
matrices,  30-32,  51-58 

adjoint,  61 

cofactor,  57 

cofactor  expansion,  57 

determinant,  56 

elementary  row  operations,  60 

Hilbert,  79 

identity,  59,  177 

inverse,  61 

invertible,  61,  177 

matrix  product,  53 
properties,  54 

minor,  57 

nilpotent,  79 

nonsingular,  61 

scalar  multiplication,  53 
properties,  53 

symmetric,  267,  309 

transpose,  54 

triangular,  60 
matrix  of  partial  derivatives,  125 
maximum,  see  extrema,  263 
Maxwell's  equations,  512-518 
mean  value  theorem,  329 

for  double  integrals,  399,  526 

for  triple  integrals,  497 

general  version,  260 
mean  value  theorem  for  integrals,  260 
method  of  least  squares,  293 
minimal  surface,  142 
minimum,  see  extrema,  263 
minor,  57 
Mobius  strip,  480 
moment,  377 

first,  377,  525 

of  inertia,  384-386,  451,  526 
second,  384-386 
total,  377 
monopoles 

magnetic,  see  magnetic  monopoles, 
516 

moving  frame,  211,217 
multiple  regression,  297 

N 

n -dimensional  volume,  560 
negative  definite,  267,  309 
neighborhood,  102,  122 
nephroid,  81 
net  force,  8 


602  Index 


Newton's  method,  176-181 
Newton,  Isaac,  194 
nilpotent  matrix,  79 
nonorientable,  479 
normal  derivative,  453,  507,  520 
normal  line 

to  a  plane  curve,  175 

to  a  surface  in  space,  175 
normal  plane,  see  plane,  normal,  22 1 
normal  spherical  image,  242 
normalization,  23 
numerical  integration 

for  functions  of  one  variable,  388-391 

for  functions  of  two  variables, 
391^100 

Monte  Carlo  method,  399 

o 

octant,  xv 
1-form 

basic,  530 

general,  530 
one-one,  see  injectivity,  83 
one-sided,  see  nonorientable,  479 
onto,  see  surjectivity,  83 
open  set,  1 02 
opposite  path,  417 
orientable,  479,  543 
orientation,  420,  479,  543 

compatible,  543 

consistent,  490,  546 

induced,  490,  546 
oriented,  420,  543 
origin 

in  R2,  xv 

in  R3,  xv 
orthogonal,  21 

osculating  plane,  see  plane,  osculating, 
211,220 

Ostrogradsky's  theorem,  see  Green's 

theorem,  436 
Ostrogradsky,  Mikhail,  436 

P 

parallel  axis  theorem,  407 
parallelogram,  7 
parallelogram  law,  4 
parametric  equations 
of  a  line 

symmetric  form,  12 
parametric  equations,  10 
of  a  cycloid,  14-15 
of  an  involute,  15 
of  a  line,  9-13 

in  terms  of  a  point  and  a  direction, 
10 

in  terms  of  two  points,  12 
parametrization,  419 
parametrized  surfaces,  455-467 

definition,  455 

smooth,  460 


area  of,  464^167 
piecewise  smooth,  463 
partial  derivative,  117 
mixed,  136 

of  higher  order,  136-138 
partial  functions,  116 
partition,  xvi,  314,  338 
path,  189 

closed,  419 

endpoints,  189 

flow  line,  225 

intrinsic  quantities,  217 

nonintrinsic  quantities,  216 

nonrectifiable,  205 

opposite,  417 

rectifiable,  205 

simple,  419 

tangent  line,  191 

velocity  vector,  190 
path  independence,  439,  440 
permeability  of  free  space,  514 
permittivity  of  free  space,  512 
perpendicular,  see  orthogonal,  2 1 
perpendicular  bisector,  76 
plane 

coordinate  equation  for,  40-43 

normal,  22 1 

osculating,  211,  220 

parametric  equations  for,  43^15 

rectifying,  221 
planetary  motion,  Kepler's  laws  of, 

193-200 
polynomial,  106,  107 
position  vector,  3 
positive  definite,  267,  309 
potential,  441 

vector,  see  vector  potential,  528 
potential  energy,  297 
potential  function,  see  functions, 

potential,  223 
potential  theory,  142 
Poynting  vector  field,  522 
principal  minors,  sequence  of,  268 
principal  normal  vector,  see  vector(s), 

principal  normal,  211 
probability  density,  406 

joint,  406 
product  rule 

for  scalar- valued  functions,  135 

nonexistence  of  a  general  form,  135 

of  a  scalar-valued  and  a  vector-valued 
function,  136 
projection  of  a  vector,  21-24 
Pythagorean  theorem,  59 

Q 

quadrant,  xv 

quadratic  form,  267,  309 

negative  definite,  267,  309 

positive  definite,  267,  309 
quadric  surfaces,  93-95 


ellipsoid,  94 
elliptic  cone,  94 
elliptic  paraboloid,  94 
hyperbolic  paraboloid,  94 
hyperboloid  of  one  sheet,  94 
hyperboloid  of  two  sheets,  94-95 
quotient  rule 

for  scalar-valued  functions,  135 
nonexistence  of  a  general  form,  135 

R 

radius  of  curvature,  240 

radius  of  gyration,  385,  451,  526 

random  variables 

independent,  407 
range  of  a  function,  83 
rectifying  plane,  see  plane,  rectifying, 
221 

reparametrization,  416 
of  a  /c-manifold,  549 
of  a  manifold,  543 
of  a  surface,  477 

orientation-preserving,  416,  418,  478, 
549 

orientation-reversing,  416,  418,  478, 
549 

reparametrization  of  a  path,  205-207 

resultant  force,  8 

Riemann  sum,  xvi,  315,  338 

right-hand  rule,  28,  36 

right-handed  system,  xv 

row  vector,  52 

s 

s -coordinate  curve,  458 
saddle  point,  265 
scalar  field,  222 

scalar  line  integral,  see  line  integral, 

scalar,  409 
scalar  multiplication,  2,  4,  36 

inR2  andR3,2 

inR",49 

properties,  3 
scalar  potentials,  445-447 
scalar  product,  see  dot  product,  19 
scalars,  1 

Scherk's  surface,  142 
second  derivative  test 

for  constrained  extrema,  286-290 

for  extrema,  265-270 
section  of  a  surface,  88 
set 

boundary  of  a,  102 

closed,  102 

compact,  271 

open,  102 
simply-connected  region,  442 
Simpson's  rule 

for  functions  of  one  variable,  390 

for  functions  of  two  variables,  394 
smooth  function,  138 


Index  603 


Snell's  law  of  refraction,  308 
solid  angle,  527 
span,  543 

linear,  543 
speed,  190 
sphere,  92 
spiral  of  Cornu,  241 
Spirograph,  79 

standard  basis  vectors,  see  vector(s), 

standard  basis,  9 
standard  normal  vector,  460 
steradians,  527 
stereographic  projection,  524 
Stokes's  theorem,  431,  490^193 
generalized,  553 

and  the  fundamental  theorem  of 
calculus,  560 
implied  by  generalized  Stokes's 

theorem,  557-558 
proof  of,  500-503 
strake,  241 
subset,  xv 
surface 

minimal,  142 

of  revolution,  see  surface  of 
revolution,  525 

oriented,  481 
surface  area,  464-467 
surface  area  element,  485 
surface  integrals,  469^488 

scalar,  470,  485 

vector,  474,  486 
surface  of  revolution,  525 
surjectivity,  83 
symmetric  matrix,  267,  309 

T 

r-coordinate  curve,  458 
tangent  hyperplane,  124,  166 
tangent  line,  to  a  path,  191 
tangent  plane,  118,  164-168 
equation,  166 

parametric  equations  for,  469 
to  a  smooth  parametrized  surface,  462 
tangent  spherical  image,  242 
Taylor  polynomial,  244,  246 
first-order,  248 
higher-order,  256 
second-order,  254 
Taylor's  theorem 

in  one  variable,  244-247 
in  several  variables 

first-order  formula,  247-248 
formulas  for  polynomials  of  order 

greater  than  two,  256 
second-order  formula,  252-254 
Lagrange's  form  of  the  remainder,  257 
telegrapher's  equation,  522 
thermal  conductivity,  520 


tofu,  263,  340 

topology,  101 

torque,  34,  243,  385 

torsion,  212-214,  217-219 

torus,  202,  459 

total  differential,  249 

tractrix,  240 

transpose,  54 

trapezoidal  rule 

for  functions  of  one  variable,  388 
for  functions  of  two  variables,  392 

triangle  inequality,  5 1 

triple  integral,  338 

in  cylindrical  coordinates,  366-368 
in  spherical  coordinates,  368-370 

2-form 
basic,  531 
general,  532 

two-sided,  see  orientable,  479 

u 

underlying  manifold,  538 

underlying  surface,  455 

union  of  sets,  xv 

unit  vector,  22,  23 

utility,  301 

utility  function,  1 86 

V 

vector  field,  189,  221 
conservative,  297,  441 
curl  of,  229-231 
divergence  of,  228-229 
gradient,  223,  297,  441 
incompressible,  229 
irrotational,  23 1 
radially  symmetric,  453 
solenoidal,  229 
vector  line  integral,  see  line  integral, 

vector,  411 
vector  potential,  516,  528 
vector  product,  see  cross  product,  28 
vector  projection,  21-24 
vector  surface  integral  element,  486 
vector(s),  1 
addition,  4 
inR",49 
properties,  2 
algebraic  notion,  1-3 
angle  between,  21 

inR",49 
binormal,  212,  217 
cross  product 

properties,  29 
cross  product,  27-36 
and  determinant,  3 1 
applications,  32-35 
definition,  28 
definition 


inR2  andR3,  1 
inR",49 
difference,  4 

displacement,  see  displacement 

vector,  5 
distance,  49 
dot  product,  18-25,  36 

definition,  19 

properties,  19 
equality 

inR2  andR3,  1 

inR",  49 
geometric  notion,  3-7 
gradient,  158 
inner  product 

inR",  49 
length,  19,  49 

magnitude,  see  vector(s),  length,  19 
norm,  see  vector(s),  length,  19 
notation,  1 

position  vector,  see  position  vector,  3 
principal  normal,  211,217 
projection  of,  21-24 
scalar  multiplication,  2,  4,  36 

inR2  andR3,  2 

inR",  49 

properties,  3 
standard  basis,  9 

for  cylindrical  coordinates,  71-72 

for  spherical  coordinates,  71-72 

inR",  50 
unit,  22,  23 

unit  normal,  outward,  432 

unit  tangent,  207,  412 

zero,  2 
vectors 

standard  normal,  460 
velocity,  190 
volume,  310 

zero,  339 
volume  element 

Cartesian,  371 

general,  365,  371 

in  cylindrical  coordinates,  367,  371 
in  spherical  coordinates,  369,  371 

w 

wave  equation,  187 

wedge  product,  see  exterior  product,  534 
Whitney  umbrella,  468 
windchill,  84,  185 
work,  26,  411 

Z 

zero  area,  319 
zero  vector,  2 
zero  volume,  339 
0-form,  530